You are on page 1of 515

Econometrics and data analysis f or

developing countries

Ecanametrics and Data Analysis far Develaping Cauntries provides a


rigorous but accessible foundation to modern data analysis and econo-
metric practice. The book contains many examples and exercises with data
from developing countries, available for immediate use on the floppy disk
provided.

Distinctive features include:


• teaching regression by example using data from actual development
experiences
• a wide range of detailed information from Latín America, Africa and
South Asia
• extensive use of regression graphics as a complementary diagnostic
tool of applied regression
• opportunities for readers to gain hands-on experience in empirical
research
• hundreds of useful statistical examples from developing countries on
computer disk

Ecanametrics and Data Analysis far Develaping Cauntries is designed as


a course consisting both of lecture and of computer-assisted practica!
workshops. It is a unique resource for students and researchers in develop-
ment economics, quantitative development studies and policy analysis.
Chandan Mukherjee is the Director of the Centre for Development
Studies, Trivandrum, India. He has over twenty years' experience of
teaching quantitative methods to economics students.
Howard White is Senior Lecturer in Applied Quantitative Economics at
the Institute of Social Studies, The Hague, The Netherlands. He has
published widely on development economics and other subjects.
Marc Wuyts is Professor in Applied Quantitative Economics at the Institute
of Social Studies, The Hague, The Netherlands. He has extensive experi-
ence as a teacher in statistics, econometrics and development economics.
Priorities for development economics
Series Editor: Paul Mosley
University of Reading

Development economics deals with the most fundamental problems of


economics - poverty, famine, population growth, structural change, indus-
trialisation, debt, international finance, the relations between state and
market, the gap between rich and poor countries. Partly because of this,
its subject matter has fluctuated abruptly over time in response to polit-
ical currents in a way which sometimes causes the main issues to be
obscured; at the same time it is being constantly added to and moditied
in every developed and developing country. The present series confronts
these problems. Each contribution will begin with a dispassionate review
of the literature worldwide and will use this as a springboard to argue the
author's own original point of view. In this way the reader will both be
brought up to date with the latest advances in a particular field of study
and encounter a distinctive approach to that area.
Econometrics and
data analysis for
developing countries

Chandan Mukherjee, Howard White and


Marc Wuyts

i~ ~~o~:~~n~~~up
LONDON ANO NEW YORK
First published 1998 by Routledge
2 Park Square, Milton Park, Abingdon, Oxon OXl4 4RN
Simultaneously published in the USA and Canada
by Routledge
270 MadisonAve, New York NY 10016
Routledge is an imprint of the Taylor & Francis Group, an informa business
© 1998 Chandan Mukherjee, Howard White and Marc Wuyts
Typeset in Times by Florencetype Ltd, Stoodleigh, Devon
Ali rights reserved. No part of this book may be reprinted or reproduced or
utilised in any form or by any electronic, mechanical, or other means, now known
or hereafter invented, including photocopying and recording, or in any informa-
tion storage or retrieval system, without permission in writing from the publishers.
British Library Cataloguing in Publication Data
A catalogue record for this book is available from the British Library.
Library of Congress Cataloging in Publication Data
A catalog record for this book has been requested.

ISBN 1O: 0-415-09399-6 (hbk)


ISBN 1O: 0-415-09400-3 (pbk)
ISBN 13: 978-0-415-09399-6 (hbk)
ISBN 13: 978-0-415-09400-9 (pbk)
Contents

List of figures ix
List of tables xiii
List of boxes XVI
Preface xvii
Introduction 1
1 The purpose of this book 1
2 The approach of this book: an example 3

Part 1 Foundations of data analysis 21


1 Model specification and applied research 23
1.1 Introduction 23
1.2 Model specification and statistical inference 24
1.3 The role of data in model specification:
traditional modelling 29
1.4 The role of data in model specification:
modern approaches 32
1.5 The time dimension in data 39
1.6 Summary of main points 42
2 Modelling an average 44
2.1 Introduction 44
2.2 Kinds of averages 45
2.3 The assumptions of the model 50
2.4 The sample mean as best linear unbiased
estimator (BLUE) 53
2.5 Normality and the maximum likelihood principie 58
2.6 Inference from a sample of a normal distribution 61
2.7 Summary of main points 71
Appendix 2.1: Properties of mean and variance 73
Appendix 2.2: Standard sampling distributions 73
Vl Contents
3 Outliers, skewness and data transformations 75
3.1 Introduction 75
3.2 The least squares principie and the concept
of resistance 76
3.3 Mean-based versus order-based sample statistics 80
3.4 Detecting non-normality in data 90
3.5 Data transformations to eliminate skewness 97
3.6 Summary of main points 106

Part 11 Regression and data analysis 109


4 Data analysis and simple regression 111
4.1 Introduction 111
4.2 Modelling simple regression 112
4.3 Linear regression and the least squares principie 114
4.4 Inference from classical normal linear
regression model 120
4.5 Regression with graphics: checking the model
assumptions 124
4.6 Regression through the origin 136
4.7 Outliers, leverage and infiuence 137
4.8 Transformation towards linearity 148
4.9 Summary of main points 159
5 Partial regression: interpreting multiple regression coefficients 163
5.1 Introduction 163
5.2 The price of food and the demand for
manufactured goods in India 165
5.3 Least squares and the sample multiple regression line 173
5.4 Partial regression and partial correlation 180
5.5 The linear regression model 184
5.6 The t-test in multiple regression 192
5.7 Fragility analysis: making sense of
regression coefficients 198
5.8 Summary of main points 206
6 Model selection and misspecification in multiple regression 208
6.1 Introduction 208
6.2 Griffin's aid versus savings model: the omitted
variable bias 209
6.3 Omitted variable bias: the theory 212
6.4 Testing zero restrictions 219
6.5 Testing non-zero linear restrictions 229
6.6 Tests of parameter stability 231
6.7 The use of dummy variables 237
6.8 Summary of main points 246
Contents vu
Part 111 Analysing cross-section data 249
7 Dealing with heteroscedasticity 251
7.1 Introduction 251
7.2 Diagnostic plots: looking for heteroscedasticity 252
7.3 Testing for heteroscedasticity 256
7.4 Transformations towards homoscedasticity 264
7.5 Dealing with genuine heteroscedasticity: weighted
least squares and heteroscedastic standard errors 270
7.6 Summary of main points 277
8 Categories, counts and measurements 279
8.1 Introduction 279
8.2 Regression on a categorical variable: using
dummy variables 280
8.3 Contingency tables: association between
categorical variables 287
8.4 Partial association and interaction 293
8.5 Multiple regression on categorical variables 295
8.6 Summary of main points 298
9 Logit transformation, modelling and regression 302
9.1 Introduction 302
9.2 The logit transformation 303
9.3 Logit modelling with contingency tables 307
9.4 The linear probability model versus logit regression 313
9.5 Estimation and hypothesis testing in logit regression 320
9.6 Graphics and residual analysis in logit regression 327
9.7 Summary of main points 331

Part IV Regression with time-series data 333


10 Trends, spurious regressions and transformations
to stationarity 335
10.1 Introduction 335
10.2 Stationarity and non-stationarity 335
10.3 Random walks and spurious regression 338
10.4 Testing for stationarity 349
10.5 Transformations to stationarity 356
10.6 Summary of main points 363
Appendix 10.1: Generated DSP and TSP series for exercises 365
11 Misspecification and autocorrelation 366
11.1 Introduction 366
11.2 What is autocorrelation and why is it a problem? 366
11.3 Why do we get autocorrelation? 370
11.4 Detecting autocorrelation 379
vm Contents
11.5 What to do about autocorrelation 387
11.6 Summary of main points 390
Appendix 11.1: Derivation of variance and covariance
for AR(l) model 391
12 Cointegration and the error correction model 393
12.1 Introduction 393
12.2 What is cointegration? 393
12.3 Testing for cointegration 399
12.4 The error correction model (ECM) 406
12.5 Summary of main points 412

Part V Simultaneous equation models 413


13 Misspecification bias from single equation estimation 415
13.1 Introduction 415
13.2 Simultaneity bias in a supply and demand model 417
13.3 Simultaneity bias: the theory 422
13.4 The Granger and Sims tests for causality and
concepts of exogeneity 425
13.5 The identification problem 428
13.6 Summary of main points 434
14 Estimating simultaneous equation models 437
14.1 Introduction 437
14.2 Recursive models 437
14.3 Indirect least squares 439
14.4 Instrumental variable estimation and two-stage
least squares 442
14.5 Estimating the consumption function in a
simultaneous system 445
14.6 Full information estimation techniques 448
14.7 Summary of main points 451
Appendix A: The data sets used in this book 455
Appendix B: Statistical tables 463
References 481
Index 485
Figures

1 Histogram of the three variables 8


2 The scatter plot matrix of the variables 10
3 Histograms with transformed data 11
4 The scatter plot matrix of the transformed variables 12
5 Scatter plot of birth against infant mortality with regression
curve (regression of birth against square root of infant
mortality) 15
1.1 The elements of a statistical study 26
2.1 The demand for and recruitment of casual labour, Maputo
harbour 47
2.2 Weekly overtime payments 48
2.3 Real manufacturing GDP, Tanzania 51
2.4 Comparing the sampling and population distributions of
labour demand 55
2.5 Comparing the sampling distributions of mean and median:
demand for labour 60
2.6 Confidence intervals of sample means: demand for labour,
Maputo harbour 64
3.1 The least squares property 78
3.2 Box plots of GNP per capita (two samples of seven African
countries) 86
3.3 Comparative box plots of gender differences in life
expectancy 88
3.4 Mean versus median: GNP per capita 92
3.5 Symmetric but unusual tail: female-male life expectancy 92
3.6 Household income data 99
3.7 Log household income 99
3.8 Fourth root transformation 102
4.1 Regressing D on R 127
4.2 Exploratory band regression: D on R 128
4.3 Scatter plot with regression line: RE on RR 130
4.4 Residuals versus fitted values 130
4.5 Exploratory band regression: RE on RR 131
x List of figures
4.6 Box plot of residuals RE on RR 132
4.7 Residuals against time 133
4.8 Regression through the origin: RE on RR 137
4.9 Graph for exercise 4.5 (Plots 1-4) 140
4.10 Studentised residual RE 143
4.11 Hat statistics versus RR 144
4.12 Box plot of DFBETA statistics 146
4.13 Energy consumption versus GNP: per capita 151
4.14 E on Y versus log(E) on log(Y) 153
4.15 Energy consumption versus urban population as
percentage of total polulation 155
4.16 E on U versus log(E) on U 156
4.17 Life expectancy against GNP per capita 157
4.18 L on Y versus L on log(Y) 158
4.19 Graph for exercise 4.15 162
5.1 Manufacturing demand versus relative food price 167
5.2 Partial regression plot of demand on food price 170
5.3 Scatter plot matrix with untransformed data 188
5.4 Scatter plot matrix with transformed data 189
5.5 Partial regression plots of regression of birth on: log income
GNP per capita; (infant mortality)º 5 ; HDI 2; and urban 191
5.6 Variation in regression coefficients across specifications 201
6.1 Prívate versus public investment: Sri Lanka, 1970-89 223
6.2 Scatter plot of S/Y against A/Y 226
6.3 Partial regression: SIY on l!Y 227
6.4 Scatter plot of life expectancy against log(GNPc) 232
6.5 Terms of trade of developing countries over time 235
6.6 Fitted line of terms of trade with intercept dummy 238
6.7 Regressions of RE on RR with and without data point 1980 239
6.8 Scatter plot of expenditure against revenue 242
6.9 Sri Lankan investment function with intercept dummy only 243
6.10 Sri Lankan investment function with slope dummy only 243
6.11 Sri Lankan investment function with intercept and slope
dummies 244
7.1 Residual versus predicted plot 254
7.2 Squared residuals versus predicted 255
7.3 Absolute residuals versus predicted 255
7.4 Log 1e1 versus predicted Y 266
7.5 Scatter plot of absolute residuals versus predicted 267
7.6 Food and total expenditure 268
7.7 Effect of log transformation 268
7.8 Residual versus predicted plot 274
7.9 Absolute residuals: regression of Fon T 274
7.10 Absolute residuals: regression of FIT on l!T 275
7.11 Comparing double-log and linear fitted regressions 275
List of figures Xl

9.1 Estimated probabilities (proportions) 305


9.2 Comparing logits 306
9.3 Scatter plot of F on A with regression line 315
9.4 Age distribution by worker category 317
9.5 Regression of O on age: linear probability model 318
9.6 Linear probability model and logit regression: Maputo
worker data 318
9.7 Linear probability model and logit regression: Indian
worker data 320
9.8 Conditional effects plots 328
9.9 Poorness of fit plot for logit regression 330
10.1 Tanzanian consumer price index, 1966-92 336
10.2 Tanzanian real export growth, 1970-88 337
10.3 Infiation and money supply: scatter plot of CPI against
M2, 1966-92 339
10.4 AR(l) process: 13 1 = O; 132 = O 339
10.5 AR(l) process: 13 1 = O; 13 2 = 0.5 340
10.6 AR(l) process: 13 1 = O; 13 2 = 1: random walk 341
10.7 AR(l) process: 13 1 = O; 13 2 = 1.1 342
10.8 AR(l) process: 13 1 = O; 13 2 = -0.5 342
10.9 AR(l) process: 13 1 = 2; 13 2 = 0.3 343
10.10 AR(l) process: 13 1 = 0.2; 132 = 1: random walk with drift 344
10.11 Two random walks 347
10.12 Trend with stationary autoregressive component 350
10.13 First differences in logged Tanzanian CPI and M2, 1967-92 351
10.14 Decision tree for testing stationarity 353
10.15 Box plot for successive sub-samples of untransformed
CPI data 357
10.16 Logged IQR against logged median for Tanzanian CPI 358
10.17 Logged Tanzanian CPI, 1970-88 359
11.1 No autocorrelation (p = O) 367
11.2 Positive autocorrelation (p = 0.7) 368
11.3 Negative autocorrelation (p = -0.7) 369
11.4 Residuals from crop production function 371
11.5 Population of Belize and fitted line from linear regression 373
11.6 Residuals from regressions of population of Belize on time 374
11.7 Residuals from crop production function from expanded
regression 376
11.8 Residuals from regression of life expectancy on income per
capita, alphabetical ordering 377
11.9 Residuals from regression of life expectancy on income per
capita, income ordering 378
11.10 Scatter plot and fitted line for regression of life expectancy
on income per capita 378
11.11 Correlogram for AR(l) process with p = O 380
xii List of figures
11.12 Correlogram for AR(l) process with p = 0.7 381
11.13 Correlogram for AR(l) process with p = -0.7 381
11.14 Correlogram for crop production function (regression 1) 382
11.15 Correlogram for crop production function (regression 2) 382
11.16 Interpretation of the DW statistic 385
12.l Actual and fitted values of spurious regression model 395
12.2 Actual and fitted values of true regression model 395
12.3 Residuals from spurious regression 397
12.4 Residuals from estimation of true model 398
12.5 Pakistani exports and real exchange rate, 1966-92 400
12.6 Scatter plot of exports and RER 401
12.7 Costa Rican consumption function: actual and fitted values 403
12.8 Scatter plot for Costa Rican consumption against income 404
12.9 Fitted and actual values for consumption function 405
12.10 Simulation of error correction model (rapid adjustment) 409
12.11 Simulation of error correction model (slow adjustment) 410
13.l The identification of the supply curve by the demand curve 429
13.2 The identification of the demand curve by the supply curve 430
Tables

2.1 Averaging the demand for and recruitment of labour at


Maputo harbour 47
2.2 Averaging overtime payments 49
2.3 Means and standard deviations: sampling versus population
distribution 56
2.4 Types of errors in hypothesis testing 67
2.5 Samples of daily demand for labour 69
3.1 GNP per capita: seven African economies, 1990 77
3.2 Resistance of mean versus median 79
3.3 Difference (female minus male) in life expectancy by level
of wealth (GNP per capita groupings), 99 developing
countries, 1990 88
3.4 The shrinking effect of logarithms 98
3.5 Summary statistics of per capita household income 100
3.6 Ladder of powers to reduce skewness 101
3.7 Fourth root transformation of per capita household
income data 101
3.8 Moment-based characteristics of a distribution 106
4.1 Share of food in total expenditure 113
4.2 Numerical summaries of the residuals 133
4.3 Plot summary 139
4.4 Summary measures of outliers, infiuence and leverage 147
5.1 All regressions with four regressors: the determinants of
the total fertility rate (TFR) 200
5.2 Minimum and maximum bounds of estimated regression
coefficient 201
5.3 Conditional bounds of coefficients of FP and CM: FP
included 203
6.1 Results of growth regressions 234
6.2 Regressions with and without intercept dummy for 1980 239
7.1 Calculation of Bartlett's test for income data 258
7.2 Summary of tests for heteroscedasticity with values for
various regressions 263
XIV List of tables
7.3 Grouped household expenditure 271
8.1 Frequency distributions of a categorical variable and a
measurement variable 281
8.2 Distribution of weekly wage earnings (per cent by gender
in each income group) 282
8.3 Regression of loge(W) on D 285
8.4 Categories of education 286
8.5 Regression of log(W) on educational levels 286
8.6 Contingency table of gender education 288
8.7 Computation of contingency chi-square statistic 291
8.8 Partial associations of weekly wage income and education 293
8.9 Average log wage income by gender and by education 294
8.10 Breakdown of average log income into components 295
8.11 Regressing log income on gender and education 296
8.12 Regression of log income on gender, education and age 297
9.1 High profit enterprises in Tanzania's informal sector, 1991 303
9.2 Tanzanian informal sector enterprises by sector, location
and age 307
9.3 Informal sector enterprises by trade/non-trade, age and
location 309
9.4 Logits in favour of trade by age and location 309
9.5 Logit estimates 322
9.6 Logit estimates: restricted version of model 325
10.1 Series resulting from different values of ¡3 1 and ¡32 in
AR(l) model 345
10.2 Frequency distribution of absolute t-statistic from 120
regressions of unrelated random walks 348
10.3 Application of decision tree for testing stationarity 354
10.4 Medians and interquartile ranges 359
11.1 Hypothetical data set for crop production function 372
11.2 Regression results for regression of population of Belize
on a trend 374
11.3 Results from crop production function regression 376
11.4 Application of Cochrane-Orcutt correction to crop
production function data 389
11.5 Regression results with Cochrane-Orcutt procedure 389
12.1 Decision tree applied to exports, RER and residuals from
cointegrating regression 402
12.2 Average and marginal propensities for consumption
function 411
13.1 Estimation of supply and demand model for Maltese
exports 421
13.2 Granger and Sims test for Maltese export quantity and
prices 426
13.3 Coefficient matrix for supply and demand model 431
List of tables xv

13.4 Coefficient matrix and rank and arder conditions for


Keynesian macro model 433
14.1 Estimates for two-commodity supply and demand model 449
14.2 Cross-products for SUR estimation 450
14.3 Estimates of simple Keynesian model for Indonesia 451
Boxes

2.1 Step-by-step procedure of hypothesis testing 68


3.1 Computing the median and quartiles 84
3.2 The definition of an outlier 87
4.1 Performing the t-test in regression analysis 135
4.2 A lognormal variable 150
5.1 Money illusion in the demand function 166
5.2 A note on notation 174
5.3 The properties of the least square line 178
5.4 The concept of auxiliary regressions 182
5.5 Counting the number of alternative specifications 200
6.1 Aid in the Harrod-Domer Model 210
6.2 The F-distribution 221
6.3 Constant returns to scale 230
7.1 White's test as RESET (regression specification error test) 262
8.1 Dummies and the constant term 284
8.2 Stochastic independence 288
9.1 Heteroscedasticity in the linear probability model 316
10.1 Logged differences and rates of growth 361
12.1 The error correction modelas a restricted time series model 408
13.1 The terminology of simultaneous systems 416
13.2 Probability limits (plims) 423
Preface

This book grew out of our frustration as teachers of data analysis and
econometrics to post-graduate students in development economics and in
population and development at the Centre of Development Studies (CDS,
Trivandrum, India) and the Institute of Social Studies (ISS, The Hague,
The Netherlands). Our main aim in both institutions was to develop a
course which puts the emphasis squarely on data analysis and economet-
rics as a research tool in the analysis of development issues. But while
many good texts exist on statistical and econometric theory, only a few
of them deal explicitly with the practice of data analysis in research, and
hardly any do so with data relating to the specific problems of developing
countries. The purpose of this book is to fill this gap.
This book would not have come about but for the successive cohorts
of students at both CDS and ISS who sat through lectures and computer-
assisted workshops based upan the successive drafts which accompanied
its development. They provided us with both encouragement and invalu-
able feedback needed to develop a book of this nature. This feedback
was particularly important to improve the design of the exercises with
real data which constitute an important element of this book. Our sincere
thanks to all these students. The development of this book also benefited
from the involvement of two of its authors (during 1991-2) in the design
and write up of the courses in econometrics and in research methods
for the MSc in Financia! Economics, the externa! programme in eco-
nomics of the School of Oriental and African Studies. Sorne of the
materials and exercises found in this book were initially developed for
this programme. Our thanks go to SOAS (and to SIDA, the funding
agency) for giving us this opportunity to participate in the develop-
ment of long-distance courses which included both conventional study
materials and computer-assisted exercises. The feedback from course
readers, tutors and students was of great help in the subsequent devel-
opment of this book.
Its writing was made possible by the clase collaboration between the
Population and Development programmes and CDS and ISS within
the framework of the Global Planning Programme in Population and
xviii Preface
Development of the UNFPA. ÜUT thanks to UNFPA for creating the
opportunity and facility for this collaboration between the two institutions.
We are grateful to Lucia Hanmer and Niek de Jong, ISS colleagues,
and to PhD students Hari Kurup, SUTesh Babu and Saikat Sinha at CDS
and Philomen Harrison and Alemayehu Geda Fole at ISS for their valu-
able comments on various drafts of the book and their willingness to check
for errors and inconsistencies. We thank Philomen Harrison also for her
assistance in setting up sorne of the data sets used here. Furthermore, the
comments of three anonymous reviewers and of Professor N. Krishnaji
(Centre for Economic and Social Studies, Hyderabad, India) were much
appreciated as they greatly helped us to improve the final version of the
book. Thanks also to Paul Mosley, the series editor, and to Alisan Kirk,
the economics editor of Routledge, for their patience and advice dUTing
the gestation period. Finally, we would like to express OUT special appre-
ciation and thanks to Annamarie Voorvelt at ISS who worked tirelessly
to turn OUT various drafts and notes into a finished manuscript.
Centre far Development Studies, Trivandrum CHANDAN MUKHERJEE
Institute of Social Studies, The Bague HowARD WHITE
MARCWUYTS
lntroduction

THE PURPOSE OF THIS BOOK


This book provides an introduction to data analysis in the context of
econometric practice. More specifically, this book deals with regression
analysis as a research tool in development studies. It is a second course
in statistics and data analysis for economists and other social scientists
working on the problems of the economy and society of developing coun-
tries. We assume that you already have done at least one basic course in
statistics for economists or for social scientists, and possibly a basic course
in econometrics as well. Hence, in this book we assume a basic familiarity
with descriptive statistics, with probability theory, random variables and
probability distributions, with statistical inference (estimation and hypoth-
esis testing) in the univariate case, and with correlation and regression
analysis, bivariate and multiple.
This book is not meant to be a specialised text for econometricians.
Instead, it seeks to equip students in development economics in partic-
ular, and in development studies in general, with a salid but accessible
foundation in the practice of regression analysis in empirical research.
This is what lends this book the following distinctive characteristics:
1 it teaches regression analysis by example by making ample use of
examples and exercises with real data (available in data files on
diskette) drawn from the experience of developing countries;
2 it pays extensive attention not just to statistical inference, but espe-
cially to model specification in applied regression analysis - a problem
of main concern to applied researchers who are often at a loss as to
which model best suits the concrete problem at hand;
3 it draws upan modern approaches to data analysis: exploratory data
analysis, misspecification testing, specification searches, and model
selection through general to specific modelling;
4 it makes extensive use of graphical methods in data analysis which,
along with hypothesis testing, constitute powerful diagnostic tools of
applied regression analysis;
2 Econometrics far developing countries
5 it deals with specific problems associated with the analysis of survey
data (heteroscedasticity; the analysis with categorical variables; logit
regression) and with regression analysis of time series data (deter-
ministic versus stochastic trends; cointegration; error correction
models).
6 it discusses the simultaneous equation bias, shows how to test for its
presence, and deals with the problem of the identification and esti-
mation of simultaneous equation models.
In dealing with these topics we choose to be selective and not encyclo-
pedic in scope. That is, while the book covers a wide range of techniques
and methods, it is not our aim here to provide an encyclopedic introduc-
tion to all the latest techniques in the practice of econometrics and data
analysis. Instead, this book puts the emphasis on explaining the underlying
principles of data analysis in applied research, and shows with a selection
of specific methods and techniques how these principles inform the prac-
tice of data analysis and econometrics without, however, being exhaustive
in coverage.
The book is an intermediate text which can be used to teach the prac-
tice of econometrics and data analysis to final-year undergraduates or to
master-level students. Its mathematical level is not high because the book
is not concerned with mathematical derivations and formal proofs of statis-
tical and econometric theory. There are many excellent introductory and
intermediate textbooks on statistics for economists and on econometric
theory at different levels of mathematical threshold. It is not the purpose
of this book to substitute for such texts. It can nevertheless be read and
studied independently since it reviews the basic assumptions and proper-
ties of classical regression analysis; most statistical and econometric results,
however, are stated without proof.
This <loes not mean that this book relegates statistical and econometric
theory to a secondary position. On the contrary, the argument of the book
is that theory matters a great <leal in applied work. It argues that it is
not sufficient to know that certain estimators or tests possess desirable
statistical properties given a set of assumptions, since, in applied work,
assumptions can never be taken for granted. It is important, therefore, to
verify whether they are reasonably satisfied in practice and to know what
to do if this is not the case. This book aims to introduce you to the prac-
tice of applying statistical and econometric methods in a world of messy
data where nothing can be taken for granted.
In statistical theory, we assume that the statistical model is correctly
specified and subsequently derive the properties of the estimators and of
the hypothesis tests based upon them. The applied researcher, however,
more often than not is mainly concerned with finding an appropriate
model for the problem at hand. A major theme running through this book,
therefore, is that, in applied research, it is important not just to test ideas
against data, but also to get ideas from data. Data help you to confirm
lntroduction 3
(or falsify) ideas you hold, but they often also provide clues and hints
which point towards other, perhaps more powerful, ideas. Indeed, applied
workers are often as much concerned with model creation as with model
estimation.
To get ideas from data it is important to look at the data. This is the
reason why this book draws heavily upon modern data analysis (in partic-
ular, exploratory data analysis) and makes extensive use of modern
graphical methods in regression analysis. Indeed, the most distinctive char-
acteristic of statistical data is their variability. No summary statistic such
as an average or a regression line can do justice to the full range of vari-
ation in data. Graphical methods allow us to look at patterns within the
data without ignoring its total variation and, in particular, the exceptions
to the rule. In fact, 'exceptions are the rule' (Levine, 1993), or, as Stephen
Jay Gould (1996: 40) put it, 'we are still suffering from a legacy as old as
Plato, a tendency to abstract a single ideal or average as the "essence"
of a system, and to devalue or ignore, variation among the individuals
that constitute the whole population'. Graphical methods allow us to see
our abstractions in the context of the full range of variation, and to assess
whether further clues and hints can be gleaned from the data which may
suggest novel ideas or things we overlooked in our analysis.
This book, we hope, may help to guard against the mechanical use of
regression analysis in applied research where a model is simply imposed
on the data without checking whether its assumptions are reasonably valid
in practice or whether it is the most appropriate model for the problem
at hand. It also guards against the all too common practice of running a
string of regressions in the hope of hitting upon one which looks good
and, hence, is singled out as the preferred model while the other results
find their way into the waste basket. To do so is data mining, a practice
rightly vilified in the literature. Specification searches and data analysis
employ similar techniques, but with a very different philosophy. Good
data analysis, we argue, involves a theory-inspired dialogue in which data
play an active part in the process of arriving at an appropriate specifica-
tion of a model, and not just in its testing. It is no use to force the data
into a straitjacket and ignore the signs which tell us that our model is
clearly inadequate, nor should we just discard bad results in the waste
basket without asking the obvious question whether the data can help
us find out why the results obtained ran counter to our expectations. In
the next section we illustrate these points with a simple example of a
multiple regression.

THE APPROACH OF THIS BOOK: AN EXAMPLE


This section uses an example to show that seemingly good results in regres-
sion analysis may turn out to be quite questionable if we care to scrutinise
our data in greater depth. It shows that the mechanical application
4 Econometrics far developing countries
of regression analysis can lead us astray. And it further illustrates how
graphical methods in particular provide us with powerful diagnostic tools
which tell us what is wrong with the model and often guide us towards a
better way of modelling our data. The example, therefore, aims to give
you an initial fiavour of the approach to data analysis we seek to develop
in this book.
It is not necessary for you to be familiar with all the techniques or
graphs used in this example to grasp the main points it makes concerning
the approach adopted in this book. In subsequent chapters we shall deal
in detail with the methods illustrated in this example. To keep the example
simple, however, we shall confine our analysis to a three-variable regres-
sion model. In developing this example we shall first adopt a more
mechanical approach to regression analysis; subsequently, we shall take a
closer look at the seemingly good results obtained with this regression;
and finally, we shall explore how the clues and hints obtained from data
analysis allow us to arrive at a better model.

Initial model specification and estimation, with sorne diagnostic testing


Suppose, then, that we intend to explain the variations in the birth rate
across countries in terms of the variations in the gross national product
per capita and in infant mortality. The birth rate of a country measures
the ratio of the number of births in ayear over the population in the mid-
year, expressed per 1,000 population. It is a fairly crude measure of fertility
since it relates total births to the total population without taking account
of the age and sex structure of the population. GNP per capita measures
the value added per capita accruing to residents of a particular country.
Infant mortality is an (in verse) indicator of the health of a nation and
measures the ratio of the number of deaths of infants under age one in
a year over the number of live births in that year, expressed per 1,000
live births.
Why would fertility depend on these two variables? Fertility may
depend on the level of income per capita for a variety of reasons (see,
for example, Dasgupta, 1993: 343-76). The conventional argument is that
the ratio of costs and benefits of children vary with the level of develop-
ment of a country. Bearing and rearing children costs time (which falls
predominantly on women) and material resources. In many agrarian
societies these costs are often spread among kith and kin, while in indus-
trialised and urbanised societies these costs mainly fall on the nuclear
family or a single mother. The costs of education, both in terms of time
and material resources, varies widely between countries, and tends to be
much higher in higher-income countries. Conversely, in terms of benefits,
in poor countries children often matter a great deal in terms of the income
they generate through paid labour or the contribution they render to the
household through unpaid labour, and as potential providers of old-age
Introduction 5
security for their parents. Taking all these factors, and others, into account,
it is usually argued that, ceteris paribus, there would appear to be a
negative relation between fertility and income (Dasgupta, 1993: 345). In
contrast, we would expect fertility to vary positively with infant mortality.
Under conditions of poverty and ill-health, where infant mortality is high,
it is the expected number of sUTviving children - that is, children who
sUTvive into adulthood - which guides fertility decisions.
Assuming the regression model to be linear, we arrive at the following
specification of the regression model (using convenient abbreviations to
denote the three variables in play):
Birth.l = a 1 + a 2 Y l¡ + a 3 IMR.l¡ + El

where birth is the birth rate, Y income per capita, IMR the infant mortality
rate, i = 1, ... , n, where n is the sample size. The error terms, the E¡S
in the model, are assumed to be each normally distributed with zero
mean and constant variance, and to have zero covariances. Following OUT
discussion above, OUT a priori expectations as to the signs of the coeffi-
cients are as follows: a 2 < O and a 3 > O. Hence, the slope coefficient of
the income variable is expected to be negative and that of infant mortality
to be positive.
Having specified OUT statistical model, we can now proceed with its
estimation. To do this, we use a sample of observations for 109 countries
in the year 1985, taken from the World Bank tables. The data can be
found in the file BIRTH on the data diskette which accompanies this
book. The least squares estimators of the regression model yield the
following results (t-statistics in brackets ):
Bifth; = 18.8 - 0.00039 Y; + 0.22 IMR; R 2 = 0.79
(11.65) (-2.23) (14.04) n = 109
At first sight, these results look good. The coefficient of determination,
R 2 , tells us that the regression explains 79 per cent of the total variation
in the crude birth rate. This is a good result given that we are working
with cross-section data and a relatively large sample size. Moreover,
both slope coefficients have the expected sign and are statistically signif-
icant at 5 per cent significance level. The near zero value of the slope
coefficient of GNP per capita should not worry you. Why? The reason
that this coefficient tums out to be so small is due to the fact that GNP
per capita, measUTed in dollars, varies over a much wider range than the
crude birth rate, measUTed as the number of births per 1,000 population.
Consequently, the slope coefficient which measUTes the impact of a $1
change in GNP per capita on the crude birth rate is bound to be very
small, but its overall effect is nevertheless substantive because of the large
variations in GNP per capita across countries.
Given that the evidence obtained from the regression confirms OUT ini-
tial hypothesis, many researchers may be inclined to stop the data analysis
6 Econometrics for developing countries
at this point. This is unwise, however, since the statistical properties of
the regression results depend on the assumptions of the regression model
being reasonably satisfied in practice. This calls for diagnostic testing of the
validity of the assumptions of the regression model. In our example, of par-
ticular importance is the normality assumption of the error term which
underscores statistical inference (e.g. the t- or F-statistics), and, since we
are dealing with cross-section data, the assumption that the error term has
a constant variance.
To test whether the normality assumption is reasonably satisfied in prac-
tice we use Jarque-Bera's skewness-kurtosis test (which will be explained
in detail in Chapter 3). The basic idea behind this test is to verify whether
the higher moments of an empirical distribution conform with those that
would be obtained if the distribution were normal. A normal distribution
has zero skew and a kurtosis equal to 3. The Jarque-Bera statistic, there-
fore, tests whether we can accept the joint null-hypothesis that a given
empirical distribution (of, say, the residuals of a regression) has zero skew
and kurtosis equal to 3. This implies the use of a chi-square test with two
degrees of freedom. In this example, the probability value of this test
applied to the residuals of our multiple regression yields a value of 0.0574,
or 5.74 per cent. At 5 per cent significance level, therefore, we would
accept the hypothesis that the residuals are drawn from a normal distri-
bution, although we should not overlook the fact that the probability value
is only just above the cut-off point of 5 per cent.
To test whether the error terms have a constant variance we shall use
the Goldfeld-Quandt test (which will be explained in detail in Chapter 7).
The idea behind this test is fairly simple. The practice of data analysis shows
that the variance of the error term often tends to increase (decrease) with
the level of one (or more) of the explanatory variables. That is, for exam-
ple, in a regression of Y on X, the variance of the error term increases
(decreases) as X increases. This is particularly common with cross-section
data. The test involves sorting the data in ascending order of each explana-
tory variable in turn, and running two separate regressions with equal
sample size for the lower and upper parts of the data, while deleting about
a quarter of the observations in the middle. If the error variance is
homoscedastic (that is, equal variances prevail), the sums of squared resid-
uals of both sets of residuals will be roughly equal in size. The relevant test
features an F-statistic which involves the ratio of the larger to the lower
sums of squared residuals of both regressions. If we do this test for our
example by first sorting the data by GNP per capita and subsequently by
infant mortality, we find that we can accept the null-hypothesis that the
error variance is homoscedastic in both cases.
This preliminary analysis may lead us to conclude that the available
evidence supports our initial model. Indeed, the coefficients of the model
have the expected signs and are statistically significant, and the coefficient
of determination indicates that the regression explains 79 per cent of the
Introduction 7
total variation in the birth rate across countries. Moreover, subsequent
diagnostic testing shows that the normality assumption is acceptable (at the
5 per cent significance level) and that the error terms are homoscedastic.
Our results, therefare, give strong evidence far our hypothesis. But how
sound are they really?

Exploring data with graphical methods


Did you notice that we never actually looked at the data when arriving
at these results? Similarly, did it strike you that the main purpose in the
analysis above was to check whether the evidence supports our hypoth-
esis without much concern about whether there is more that could be
learned from the data? The argument of this book is that data analysis
can be made more productive and more lively if we care to look at our
data. But how do we do this?
The first important lesson of data analysis is that, as Hamilton (1992:
1-2) put it, it is always preferable to proceed from the ground up by first
carefully study each variable in turn, and then in pairs, befare proceeding
to more complex analyses. This is the main reason why this book features
two chapters on univariate analysis befare moving on to regression
analysis. But why is this so important?
The main aim of regression analysis is to explain the total variation in
a dependent variable by breaking it down into the explained variation
due to the explanatory variables included in the model and the residual
variation. It is useful, therefare, to have a good preliminary look at the
nature of the variation (of the dependent variable) to be explained and
to compare it with the character of the variations in the explanatory vari-
ables. To do this, we look at the empirical distribution of the data of each
of the variables in play. Figure 1 uses histograms to capture the variation
in each of the variables in our model.
The most striking feature about Figure 1 is that the three histograms
differ radically in shape. While the crude birth rate appears to be vaguely
rectangular in shape, GNP per capita is strongly skewed to the right, and
infant mortality is moderately skewed to the right. In general, regressions
between variables which differ significantly in shape do not perfarm well.
Ideally, with non-experimental data, we would like our variables to be
similar in shape. Why? The econometrician Granger (1990: 12) gives us
the fallowing important principie as a possible reason: 'if a variable being
explained by a model, the dependent variable, has sorne dominant
features, a minimum condition far the model to be satisfied is that the
explanatory variables should be capable of explaining this feature.' And,
undoubtedly, the shape of the distribution of the dependent variable is
one of its most important features, and, hence, it is important to verify
the shapes of the distributions of the variables in the model. Furthermore,
regression theory suggests that we would like our variables to be at least
0.146789 0.587i56

o --~~~~~~~~~~~~~~~~~~ o '-r--~~~~-.--~~~--,r--~~~~~~~~~

10 54 110 Per capita GNP


19270
Crude birth rate per 1000 pop.

0.211009

o --~~~~~~~~~~~~~~~~~~
6 lnfant mortality (age<1 year)
175

Figure 1 Histogram of the three variables


Introduction 9
more or less symmetrical in shape, preferably bell-shaped. In our example,
this is not the case. This brief glance, therefore, at the univariate distrib-
utions of the data gives us a warning that regression analysis with these
data may not be without problems.
Let us go one step further, while still proceeding from the ground
upwards, by looking at the bivariate relations between the variables in
play. This can best be done by using the simple but powerful device of
the scatter plot matrix as shown in Figure 2.
The scatter plot matrix is the graphical equivalent of a correlation matrix
and features the pairwise scatter plots of the variables included in a
multiple regression. Hence, for example, the plot in the first row and
second column features the scatter plot of GNP per capita against infant
mortality, while the plot in the second row and first column shows the
same plot with the axes reversed (infant mortality against GNP per capita),
and so on. The diagonal plots are omitted because they merely involve
plotting a variable against itself. The third row features the plots of the
dependent variable against each of the explanatory variables.
All the plots are non-linear in shape. This is particularly clear in the
plot of the birth rate against GNP per capita (third row, first column) and
in the plot of infant mortality against GNP per capita (second row, first
column), but also, to a lesser extent, in the plot of the birth rate against
infant mortality (third row, second column). This should not surprise you,
given the marked skewness in the distributions of the explanatory vari-
ables as distinct from that of the dependent variable. Furthermore, sorne
of the plots also reveal the presence of outliers: points which do not fit
the general pattems inherent in their scatters.
But undoubtedly the main features of the data are the pervasive pres-
ence of skewness in the explanatory variables and the presence of non-
linearities in the pairwise scatter plots. As we shall show in subsequent
chapters, this calls for appropriate transformations of the explanatory vari-
ables aimed at eliminating the effects of the skewness in our data. Not
uncommonly, this type of transformation also helps us to correct the
non-linearities in the scatter plots. The application of Tukey's ladder of
transformations (explained in Chapter 3) led us to the decision to use
the logarithms of GNP per capita and the square root of infant mortality,
while leaving the birth rate unchanged. Figure 3 shows the corresponding
histograms of the variables thus transformed.
As can be seen from these histograms, the shape of the distributions
of the new variables, while still different, are now quite similar as far as
symmetry is concemed. There is no longer any pronounced skewness in
any of these distributions. In fact, the empirical distributions above all
appear more rectangular in shape.
Figure 4 shows the corresponding scatter plot matrix. It can be seen
from this that the non-linear pattems which dominated the scatter plot
matrix with the raw data (depicted in Figure 2) have largely disappeared.
6 175
o o 19270
r:P o o
8 o o o o

~ t;f8o
Per capita GNP
~ o ºº
éP o
o
~o o o o r:P o cP
a=' o o
o o 8

a9~ ES~~o
0

~~~8'~ttioo
000
ofb'bº 110
175 ~ oº
a> O::P o

lro
<bcPºº
o o o o o
o oº~gq,
lnfant mortafity '8€1 cfbº o
o 0
(age<1 year) o0 o 'b 'bºº
o co oG::io 0º
o
o cf' ºº o

6
'&f§ o
ºo.9 c9:tl %~ ~ 00
o
6l
~~:d>t o
o o

~oº ~o
54
o o Oc> 0
g
0€ 0
0 08 EP4ºo ( º o
o ºoº<b o ºo

~
o o 'gorn o Crude birth rate
o o o o 08 aD ~00oa:o
o o
0 per 1000 pop.
«Jf> oºº

~~
ºº o
a't8 o
o & o o o
8 00
~ 88 ~o o o 10
110 19270 10 54

Figure 2 The scatter plot matrix of the variables


0.146789 0.119266

o '-T-~~~--,~~~~-.-~~~~,.--~~~--.- o L.-~~~~..----~~~---..~~~~---,-~~~~.-
10 Crude birth ra1e per 1000 pop.
54 4. 70048
Log GNP per capita 9.8663

o '-T-~~~~.,...-~~~---.-~~~~-.-~~~~~

2.44949 13.2288
Square root infant mortality

Figure 3 Histograms with transformed data


2.44949 13.2288
o o
9.8663
i:
8 ºo o
o

o
o o
9J oº
cP~
o ocP
8
0
o0
o
o
o cP

o
Lag GNP per <!lb0 o o o o ººº o o o º'b o

~O
Cl> ieébo 00 O O O o8o<fJO Offi o
cap ita 8
o~ º8,'.lh 8 o
QOO @ o0 o o... o º ..cP cPo C!-i
o o " & EJ"" roe>&&
8 §0 8o"u~ o 0 o o ºo~~
o o oooo °
0
o o o ºªo ro
4.70048
13.2288 o o oº
o o cO 8
Cb~'Wo
0

~oº@º%:>
~~ººo&, ºº
o
o o 93 €1
o 0
oº g&:, o
o
O ~ Q)O Square root co8 :;,_ 'b o_n
0 -vO 00 \T
o o
o cJ3 o o
o
o (JI) o o infant mortality 8 o
ºº o oi ºº ó) o o
oº&º~
0
o o oo>º Ef>
~o o o
0o 0
ºº~-º
co~~<:!IO
o
2.44949
o o o o o 54
~o
€1 ~~8gt,o o o 0o
0ºº0::;º9..e
?iJ~ ¿;:? o
o ~ oº @ 6' ºo o
c9
o o 0 o0 0o --g6° ca o
o a9~ o ~ o o
<go(JI) o
Crude birth rate
& º~ºooº
o
oº o o o& Gt> sººocro
o o
o o ºo o o oº
per 1000 pop.
ocrPooO
o Sl. o o o
o oé9o
o o oCf
8 ºº
8ffÍ'j
~
0
aeci
o
¿ o o
o'~
10
4.70048 9.8663 10 54

Figure 4 The scatter plot matrix of the transformed variables


Introduction 13
Each of the scatter plots now displays a linear pattern, although sorne still
reveal the presence of outliers. More specifically, the plots featuring GNP
per capita on one of their axes contain outliers both above and below the
main pattern in the data. Por example, the plot of the birth rate against
GNP per capita shows that, on average, as GNP per capita increases, the
birth rate declines, but sorne points do not conform with this general
pattern. Sorne countries (about four) show a high birth rate given their
GNP per capita, while two countries in particular have a low birth rate
given their GNP per capita.
In data analysis, these type of clues and hints given by the data can be
invaluable to push the analysis further. In this case, for example, the two
outliers which lie significantly beneath the main pattern in the data are
China and Sri Lanka: both countries have a low birth rate given their
GNP per capita (and, similarly, both countries have a low infant mortality
rate given their GNP per capita). The countries which lie significantly
above the main pattern of the data are mainly fairly rich oil-producing
countries with high GNP per capita, high birth rates, and also high infant
mortality. These outliers show that it is not always true that a higher
(lower) income per capita implies a lower (higher) birth rate (or lower
(higher) infant mortality). They are the exceptions to the rule which can
give us deeper insights into the problem at hand. Note, however, that the
scatter plot of the birth rate against the (square root of) infant mortality
does not show the presence of any outliers. This is an important point
which can help us to model our data better.
The graphical analysis of the shape of the distributions of each of the
variables and of their pairwise scatter plots suggests that it may be prefer-
able to try out a regression with the transformed explanatory variables.
lt is to this that we turn to next.

A regression model with transformed explanatory variables


The graphical analysis above prompts us to reformulate the specification
of our model of the birth rate as follows:
Birth; = a 1 + a 2 log(Y); + a 3 ...JIMR; + E;

The estimation with least squares of this model yields the following results
(t-statistics in brackets ):
Biith; = -2.59 + 0.63 log(Y); + 4.06 ...JIMR; R 2 = 0.85
(-0.38) (0.925) (13.78) n = 109
The first thing to note about this new regression is that its coefficient
of determination is now about 85 per cent, as against 79 per cent in our
earlier regression. Since both regressions feature the same dependent vari-
able, the crude birth rate, their R 2s are comparable because they are both
ratios of the same total sum of squares (i.e. the total sums of squares
14 Econometrics far developing countries
of the crude birth rate). Note, however, that the slope coefficient of the
logarithm of GNP per capita no longer has the expected sign, nor is it
statistically significant at the 5 per cent significance level.
This lack of significance suggests that the income variable should be
dropped from the equation altogether. It is possible to check this propo-
sition graphically using a partial regression plot (as we shall show in
Chapter 5). Here we just report the regression results obtained by drop-
ping the income variable (t-statistic in brackets):
Biith = 3.61 + 3.83 -.JIMR; R2 = 0.85
(2.75) (24.17) n = 109
This simple regression confirms that dropping the income variable from
the equation hardly affects the coefficient of determination. This regres-
sion, therefore, yields a better result than the multiple regression of the
birth rate on both GNP per capita and infant mortality.
But perhaps you may be inclined to think that the loss of importance
of the income variable, GNP per capita, is solely due to the use of the
logarithmic transformation which may have been inappropriate in this
case. But the results of the following regression of the birth rate on GNP
per capita and on the square root of infant mortality (t-statistics m
brackets) shows that this is not the case:
Biith; = 3.31 + 0.00003 Y; + 3.86 -.JIMR; R2 = 0.85
(1.56) (0.185) (17.61) n = 109
Clearly, GNP per capita is not statistically significant and, hence, can
be dropped from the equation without any significant change in the coef-
ficient of determination. The simple regression of the birth rate on the
square root of infant mortality, therefore, is superior to any regression
model which also includes GNP per capita or its logarithm.
As can be seen from Figure 3, the distributions of the birth rate and
the square root of infant mortality are quite similar: both tend to be rectan-
gular in shape. Furthermore, Figure 4 shows that the scatter plot of the
crude birth rate against the square root of infant mortality (third row,
second column), unlike the other scatter plots in this matrix, does not
feature any outliers.
Further testing also shows that the residuals of the simple regression
of the birth rate on the square root of infant mortality lead us to accept
the hypothesis that they are drawn from a normal distribution. In this
case, the probability value of the Jarque-Bera skewness-kurtosis test
equals 0.165, or 16.5 per cent, which is well above the cut-off point of
5 per cent. The scatter plot of the birth rate against the root of infant
mortality reveals a very slight tendency towards heteroscedasticity, but
this is unlikely to be very significant. In fact, the application of the
Goldfeld-Quandt test leads us to accept the null-hypothesis of homoge-
neous error variances.
Introduction 15
54.3432
o
o
ro o
o o o o o
o o o o ro
o o o
o o o
ci o o 00 o o o
o o
c. o o
o o o o
o o
o o
o
o o
o o
o
00
o o
o
00 o o
o o o
o
o o o
o
o o o
00 o
o o
o
o o
o o o
o
o
o
ro o
00 o
00
10 000

6 175
lnfant mortality (age<1 year)

Figure 5 Scatter plot of birth against infant mortality with regression curve
(regression of birth against square root of infant mortality)

Figure 5 shows the scatter plot of the birth rate against infant mortality
(untransformed) along with the predicted regression curve obtained by
regressing the birth rate on the square root of infant mortality. As can
be seen from this figure, the slope of this regression curve declines as
infant mortality increases. That is, at higher levels of infant mortality it
requires a much greater reduction in infant mortality to reduce the birth
rate by a given amount than it does at lower levels of infant mortality.
Intuitively, this makes sense. If parents' decisions on fertility are deter-
mined by their concern with the number of children who survive into
adulthood, changes in infant mortality when its level is already low is
likely to have a much bigger impact on the birth rate than similar changes
when the level of infant mortality is high. Indeed, in the latter case, infant
mortality still remains high (even if it declines somewhat) and, hence, the
risk of children not surviving still remains considerable. This might explain
why the square root of infant mortality performs better as an explana-
tory variable than infant mortality itself.
This example showed that the transformation of the explanatory vari-
ables led us to adopta simpler model which no longer features GNP per
capita as an explanatory variable. This may have come as a bit of a surprise
since, clearly, the level of income would appear to be an important factor
in explaining the variations in the birth rate. Our results, however, do not
imply that the level of income does not matter at all. What they say is that
GNP per capita has nothing to add in terms of explaining the variation
16 Econometrics far developing countries
in the birth rate once we have already taken account of the influence of
infant mortality on the birth rate. But clearly the health of a nation in part
depends on its wealth in general, and on its average income in particular.
A quick glance back at the scatter plot of infant mortality against GNP per
capita (or, better still, of the square root of infant mortality against the
logarithm of GNP per capita) tells us that both are clearly related. But these
plots also reveal that sorne countries (in particular, China and Sri Lanka)
have low infant mortality despite their low GNP per capita, while other
countries (such as sorne of the richer oil-producing countries) have a high
infant mortality despite their high GNP per capita. In these exceptional
cases, the variation in the birth rate tallies with the variation in infant
mortality, and not with that in GNP per capita. Hence, health appears to
matter more than wealth in explaining fertility, yet clearly the health of a
nation depends to a great extent on its wealth.

Exercise
The data set SOCECON (available on the data diskette) which features
a set of socioeconomic data for a sample of countries for the year 1990,
contains observations for the birth rate, GNP per capita, and infant
mortality. Use these data to repeat the analysis carried in this section
(which was done with data for the year 1985) and verify whether you
obtain similar results for 1990.

Conclusion
The example in this section shows that we cannot always take our initial
model and the numerical results obtained from it at face value. Results
which look good at first sight may be riddled with problems if we care to
look at our data more carefully. Many of the problems which emerge at
the level of multivariate analysis can often be traced back to particulari-
ties of the data we encounter in studying their univariate distributions and
their pairwise scatter plots. And, as shown in this book, even when we
move to multivariate analysis it is still possible to combine diagnostic
testing with the use of various simple yet powerful graphical methods
(such as, for example, partial regression plots) which allow us to look in
depth at the results of multiple regressions. Graphical methods of data
analysis and careful diagnostic testing are indeed the principal tools which
allow data to play an active part in model specification and evaluation,
and as such are invaluable instruments in applied research.

THE STRUCTURE OF THE BOOK


The book is divided into five parts. Part I lays the foundations for the prac-
tice of regression analysis. Chapter 1 contrasts traditional and modern
Introduction 17
approaches to model specification. It argues that, while traditionally (in
theory at least) data were not supposed to play any role in model specifica-
tion, modern approaches recognise the importance of data analysis in model
creation, and not just in model estimation and testing. The chapter also
briefly discusses the specific problems associated with modelling cross-
section as against time-series data. Chapters 2 and 3 deal with the analysis
of univariate data. Chapter 2 discusses the centrality of the normality
assumption in statistical analysis and shows that if data are drawn from a
normal distribution, the mean is the most powerful estimator of the average
of the distribution. In practice, however, socioeconomic data are seldom
normal in shape. Chapter 3 goes on to see what happens if an empirical
distribution cannot reasonably be approximated by a normal distribution,
and shows that the mean, which is a least squares estimator, rapidly loses
its desirable properties. More specifically, it shows that skewness and the
prevalence of outliers in an empirical distribution often render the mean a
highly misleading average. The chapter shows how to test for normality and,
if needed, how to correct for skewness using appropriate power transfor-
mations so as to render the empirical distribution symmetrical and prefer-
ably bell-shaped. Finally, the chapter shows why it is that, with most
socioeconomic data, it is the median, rather than the mean, which provides
us with a more reliable estimator of the average of a distribution.
Part II deals with the practice of applied regression analysis and consti-
tutes the core of the book. Chapter 4 discusses simple regression, while
Chapters 5 and 6 deal with its extension to multiple regression. As a point
of departure, this part briefly reviews the assumptions and properties of
the classical linear regression model (first simple, then multiple regres-
sion) and the standard statistical inferences (estimation and hypothesis
testing) based upan it. But, as argued in the chapters of this part, we can
never take the assumptions of the model we use in practice for granted,
nor do we always know a priori whether our model is correctly specified.
The problem is further complicated by the fact that, in development
research (as in socioeconomic analysis in general), we work almost exclu-
sively with non-experimental data which are often highly collinear in
nature and, hence, render the interpretation of results difficult. The
novelty of the material discussed in this part, therefore, concerns the treat-
ment of the practice of regression analysis when it is uncertain what the
most appropriate model is and whether its assumptions are likely to be
satisfied in practice, and when the interpretation of the resulting regres-
sion coefficients is often obscured due to the prevalence of collinear
regressors in socioeconomic research. This explains the importance
devoted in this book to residual analysis and influence diagnostics; to
transformations towards normality and linearity; to the interpretation of
regression coefficients, using partial regression and partial regression plots,
in situations where nothing is held constant; to the sensitivity of regres-
sion coefficients to alternative neighbouring specifications with collinear
18 Econometrics for developing countries
regressors; to the practical problems associated with the potential trade-
off between the misspecification bias and the precision of an estimator;
and to the practice of testing down by imposing linear restrictions on a
general model to arrive at a more specific specification.
Part 111 then turns to particular problems of regression and data analysis
which are mainly (but not exclusively) associated with the use of cross-
section data. Chapter 7 deals with heteroscedasticity: how to detect it (using
both graphical methods and diagnostic tests) and how to correct for it,
either through the use of an appropriate transformation of the data or
through the use of weighted least squares. Chapter 8 deals with the use
of categorical data in regression analysis. This is particularly important in
development research where survey data (in which categories and counts
are often prominent) occupy an important role in applied research. Apart
from the use of categorical data in regression analysis by means of dummy
variables, this chapter also discusses the analysis of contingency tables
and test of independence between categorical variables. Finally, Chapter 9
looks at the problem when the dependent variable is categorical and
dichotomous in nature, and discusses the use of logit modelling and regres-
sion in applied research. Specific attention is paid in this chapter to the logit
as a convenient transformation of counts of a binary categorical variable,
and to residual analysis and influence diagnostics in logit regression.
Part IV deals with sorne of the problems associated with regression
analysis with time series data. Chapter 10 discusses why the problem of
spurious correlation is so prevalent with trended time series data. It intro-
duces the concepts of stationarity and of deterministic and stochastic
trends in time series, and shows how to test for stationarity. Chapter 11
deals with autocorrelation as a problem of model misspecification and
shows how to test for its presence and what lessons can be learned from
it in terms of model specification. Finally, Chapter 12 gives an introduc-
tion to the analysis and application of cointegration and of error correction
models in applied research.
Finally, Part V deals briefly with the implication of simultaneity in
economic models for empirical analysis. Chapter 13 discusses the problem
of the simultaneous equation bias and shows how to test for the exogeneity
of the explanatory variables in a model, and subsequently deals with the
identification of a simultaneous equation model. Chapter 14 then discusses
single equation methods for estimating an equation in a simultaneous
model.

USING THE DATA SETS WITH THIS BOOK


This book puts the emphasis squarely on data analysis in econometric
practice as applied to problems of developing countries. As stated above,
our aim in this book is not to provide specialised training in advanced
econometrics, but rather to provide applied researchers and students of
Introduction 19
development problems with a salid foundation in the practice of empir-
ical research. Our experience as teachers in econometrics, data analysis
and development has taught us that this can best be done by giving
students ample opportunity to practice with real data drawn from prob-
lems of developing countries. For this reason, the data sets used in the
majar examples and in the various exercises of this book are included in
the diskette with the book.
Appendix A, on the data sets, lists the definitions and sources of the
variables used in each of the data sets provided on diskette and used in
the examples and exercises of this book. These sets contain both cross-
section and time-series data. Sorne of the data sets contain comparative
data for a sample of countries, sorne are secondary data relating to one
country only, and the remainder feature primary survey data drawn from
micro studies of specific problems. All data sets relate to problems of devel-
oping countries, a focus which is generally not found in textbooks on econo-
metrics and data analysis. Taken together, these data sets provide teachers
with many examples and students with ample opportunity to gain hands-
on experience with real data. In the main text of this book each data set is
identified by its file name without extension. The data diskette then con-
tains the data sets both as LOTUS and ASCII files. Hence, in the text of
this book, the file BIRTH refers to the data used in the example of this
introduction; on the data diskette, the corresponding data can be found as
BIRTH.WKl in a LOTUS file andas BIRTH.TXT in an ASCII file.
To be able to carry out the exercises in this book or to redo its exam-
ples, students should have access to a computer. Many of the exercises of
the book can be done with a spreadsheet program; in teaching we find
that first introducing examples through a spreadsheet encourages students
to better understand the basic concepts. Even though most spreadsheets
do not calculate statistics such as the Durbin-Watson statistic, such output
may be fairly readily constructed. However, in more advanced applica-
tions the use of a spreadsheet can become quite tedious. Hence it is
preferable to be equipped with an appropriate statistical and/or econo-
metric software package. We found it quite useful to rely on both a
statistical package and an econometric package: the examples in this book
were worked out with either STATA or TSP, depending on which was
most convenient to use (although STATA is not so common, few other
packages have its range of data analysis tools). The use of the data sets
in this book, however, in no way depends on the use of any particular
statistical or econometric package, although obviously sorne packages lend
themselves more easily to the approach adopted here.
In general, a statistical package equipped with good analytical graph-
ical methods of modern exploratory data analysis lends itself best to
teaching regression applications, particularly when coupled with univariate
analysis. They are also useful for teaching cross-section analysis, particu-
larly when it involves the extensive use of categorical data. Econometrics
20 Econometrics far developing countries
packages, in contrast, have a clear advantage with time series data. Hence,
the analysis of stochastic or deterministic trends, the use of lagged vari-
ables in regression models, and, more specifically, the analysis of
cointegration and of error correction models, can best be done with econo-
metric software. As a general rule, however, the approach of this book
favours the use of statistical or econometric packages which are equipped
with good analytical graphics.
Part 1
Foundations of data analysis
This page intentionally left blank
1 Model specification and
applied research

1.1 INTRODUCTION

In development research, most research questions typically involve rela-


tions between two or more variables. Regression provides a powerful tool
to investigate such relations empirically. This book deals with applied
regression as a research tool. Applied regression is about modelling data,
a complex process of the evolving specification, estimation, evaluation and
interpretation of a model (Granger, 1990: 1). Good modelling requires an
active dialogue between theoretical refiection and empirical evidence
to arrive at a model which presents an adequate, yet parsimonious, approx-
imation of the underlying mechanisms helping to bring about the
phenomena to be explained.
You may find the task of modelling data rather daunting notwith-
standing the fact that you may have received prior basic training in
statistical or econometric principies. Working with real data always turns
out to be less straightforward and definitely more messy than conven-
tional textbooks make it out to be. Part of the problem is that good
modelling is both an art as well as a science and, hence, a good grounding
in statistical and econometric theory is necessary but not sufficient. But
it is equally true that traditional textbooks in statistics or econometrics
do not always address the type of concerns which are foremost in the
mind of applied researchers. They often pay undue attention to estima-
tion and hypothesis testing within the confines of a given model, but have
little to say about the difficulties involved in arriving at an appropriate
model specification in the first place. In other words, traditional texts tend
to assume that a researcher is fully equipped with a correctly specified
model which he or she seeks to estimate or test against data specifically
sampled for that purpose. In applied work, however, we cannot make this
assumption so easily. In fact, model creation or selection is often the most
creative and exciting but also frustrating aspect of doing research. The
search for an appropriate model which answers the research question of
an empirical study is more often than not the main preoccupation of an
applied researcher. But, unfortunately, traditional textbooks in statistics
24 Econometrics far developing countries
and econometrics remain rather silent on this issue. This book puts the
emphasis squarely on modelling data by applied regression analysis.
Obviously, to do this we shall make extensive use of regression theory
within the context of formal statistical inference. But, while reviewing the
basics of regression theory, our point of view will consistently be that of
an applied researcher who seeks to put to use what he or she has learned
in statistics or econometrics courses in a context where he or she is preoc-
cupied with finding out what the appropriate model is that best answers
the research question which guides empirical work. It is this viewpoint
which gives this book its particular fiavour.
This chapter addresses the question of the role that data play in model
specification. This is not an easy question nor is there agreement as to its
possible answers. It is, however, a question of great practica! significance.
The way you answer this question will shape the modelling strategies you
are likely to adopt or consider permissible, although it is not uncommon
for a researcher to profess one answer in principle, yet behave quite differ-
ently in practice. The key issue is whether data have a role to play in
model creation or whether they should only be used to test a given model
or estímate its unknown parameters. In other words, do we only test ideas
against the data or do we also get ideas from analysing data? This is a
highly contentious issue which underlies debates on modelling strategies
in applied research. Section 1.2 gives a simple framework to clarify the
connection between statistical inference (model estimation and testing),
on the one hand, and model specification, on the other. Section 1.3 then
discusses traditional econometric modelling in which data were not meant
to play any role in model specification, although in applied work ad hoc
modifications were made to improve a model. In modern parlance, as a
result of mounting criticism against this approach, this type of modelling
is known as specific-to-general modelling or the average economic regres-
sion. Section 1.4 discusses three modern approaches to modelling: general-
to-specific modelling, exploratory data analysis, and fragility analysis
(known, more technically, as extreme bounds analysis). Finally, section
1.5 briefiy discusses the distinction between cross-section and time-series
data in the light of the problems they pose in terms of modelling data.

1.2 MODEL SPECIFICATION ANO STATISTICAL INFERENCE


In development research, our interest is to investigate relations between
two or more variables. The types of relations we are talking about seldom
concern exact relationships between variables unless we are dealing with
identities in a context of accounting frameworks. Typically, our problem
is to come to grips with relations between variables in non-deterministic
situations in which regularity of data goes hand in hand with consider-
able random error fiuctuations. For example, Engel's law, which postulates
that the share of food expenditures in total household expenditure falls
Model specification and applied research 25
as household income rises, does not imply that two households with equal
income will spend an equal amount of money on food. On the contrary,
Engel's law at most refers to an average relation. Hence, most of the
relations we investigate in applied work tend to be blurred due to unac-
countable erratic variation in the data. The types of relations we deal
with, therefore, are essentially imperfect.
This is the reason why we rely on statistical modelling to analyse rela-
tions between economic variables. Typically, statistical data display two
distinctive features: a regularity of sorne kind coupled with erratic varia-
tion. A statistical model based on probability theory seeks to capture both
these components (regularity and error variation) in a single specification,
a model, of the way the data behave. This theoretical stochastic specifi-
cation, therefore, embraces both a systematic structural component (which
can be a simple average or a more complex average relation between
several variables) and a random component, which taken together account
for the total variation in the dependent variable.
The easiest way to think about this distinction between regularity and
residual random variations is to refiect on the difference between sound
and noise. If you listen to a badly tuned radio, you will have difficulties in
distilling the meaningful message (the sound) in the midst of the interfer-
ing noise. Statistical modelling tries to do something similar when analysing
variability in data. The aim is to extract a meaningful message (the sys-
tematic component) in the midst of erratic variations (the noise element).
A well-specified model, then, conveys a clear message (sound) surrounded
by unexplained and irreducible error variation (noise). What matters is not
just that the structural part of the model is theoretically meaningful and
incisive, but also that the noise component no longer contains any signifi-
cant messages. Perhaps paradoxically, therefore, to check whether an
estimated model is reasonably adequate we need to take a good look at
the residual variation it leaves unexplained to see whether it no longer
contains hidden messages which signal that the model is probably mis-
specified. Put differently, to test for model adequacy we do not just look at
the message its structural component conveys, but we also scrutinise its
residual variation to check whether it leaves things unsaid.
A statistical model, therefore, is an abstraction we use to characterise
and explain the variability in real data. It is a theoretical construction in
a double sense. First, to model the data we draw upon substantive theory
(e.g. economics, population studies, social analysis ), and second, we also
rely on probability and statistical theory to model the stochastic nature
of the relations between variables. Both sources of theoretical inspiration
- substantive theory and statistical theory - do not join together, however,
in an additive fashion. As we shall see repeatedly in this book, modelling
does not just involve tagging on a random component after the substan-
tive analysis has been completed. The choice of the random component,
the error term, is itself an important part of the modelling exercise.
26 Econometrics for developing countries

MODEL
REAL POPULATION:
POPULATION THEORETICAL
ABSTRACTIONS

/
J~
''

,J CORRECT .._
~ ~

SPECIFICATION?

,,. V

REAL MODEL
SAMPLE: SAMPLE
OBSERVE O
DATA

Figure 1.1 The elements of a statistical study


Source: Adapted from Giere, 1991: 126-8.

An empirical study involves a dialogue between theory and empirical


evidence so as to arrive at an explanation which provides an adequate
answer to our research question. This dialogue implies a confrontation
between abstract theory-inspired models and the data, in a process which
is generally interactive. Figure 1.1 provides a schematic overview based
on Giere (1991: 126-8). It has a simple but powerful message.
Giere makes a distinction between the four components of a statistical
study: the real population, the real sample, the model population and the
model sample. To start with, in statistical analysis we sample data from a
real population so as to be able to make valid inferences from the real
sample about its population. In other words, we seek to generalise our
conclusions derived from a sample beyond the confines of the sample.
To do this, however, we need to make explicit assumptions about the
character of the real population and about the nature of the sampling
procedure. This is where theoretical abstractions come into play. The
model population is a hypothetical theoretical construction (a model)
which depicts our assumptions about the character of the real population.
It is, therefore, a theoretical population which behaves exactly in accor-
dance with our assumptions. Finally, a model sample is a hypothetical
sample drawn from the model population in accordance with our assump-
tions about the sampling procedures.
Model specification and applied research 27
The theory of statistical inference allows us to make valid inferences
about the model population from the model sample. It deals with esti-
mation and hypothesis testing in a context where the data behave exactly
as our assumptions would have them behave. If our assumptions are
reasonably valid in practice, the real sample will more or less behave as
a model sample would do, and, hence, we can reasonably safely proceed
by making inferences from the real sample about its population as
characterised by the model population. Our inferences, therefore, proceed
from real sample to model population, as indicated by the diagonal line
in Figure l. l.
Hence, to be able to generalise conclusions drawn from a real sample
we always need to make explicit assumptions about the character of the
real population. But for these inferences to be valid in practice, our theo-
retical model has to be a reasonably adequate specification of the character
of the population and of the nature of the sampling procedure. Otherwise,
any conclusions we draw from our sample are invalid notwithstanding the
outward sophistication of models and techniques we use.
This is what the art of modelling is all about. It is very important, there-
fore, that we should never confuse the real data with the abstractions we
use to model them. There are two reasons for this. First, it would lead us
to accept models at face value without investigating whether they are
reasonably good approximations of the patterns inherent in the data. In
other words, as Leamer (1978: 4) put it, in applied work we cannot take
the 'axiom of correct specification' for granted. That is, we cannot just
assume that our model correctly specifies the character of a real popula-
tion and of the sampling procedures used to obtain the observed data.
Part of the exercise of modelling is the need to verify whether the assump-
tions we make are reasonably valid in practice. But, second, even if this
is the case, we should never forget that real data never behave exactly as
the theoretical variables we use to represent them.
To see this latter point, consider a simple hypothetical example of
a modelling exercise with only one variable. Suppose we have a sample
of observations on daily wages of casual workers (paid on a piece-rate
basis ), randomly drawn from a wider population of daily wages paid
out over a given period in time. As we shall see in Chapters 2 and 3, we
might consider modelling the data by assuming that they were indepen-
dently sampled from a normal distribution with unknown mean and
variance. This assumption now defines a theoretical stochastic variable
which depicts the assumed behaviour of our real data on daily wages. If
the real data display a typical bell-shaped empirical distribution (without
majar outliers), we can safely base our inferences on this assumption
and proceed, for example, by estimating the mean and variance of this
distribution using the sample data.
But daily wages cannot be negative nor can they exceed sorne defi-
nite upper limit. In contrast, a theoretically normally distributed variable
28 Econometrics for developing countries
ranges over a domain from minus to plus infinity. Given a big enough
sample, it is always possible to show that the real data diverge from our
theoretical assumptions (Chambers et al., 1983: 192). Real data can never
come from a genuine normal distribution with its exact shape, infinite
precision and range. What matters in practice, however, is whether this
theoretical bell-shaped curve provides a reasonable approximation of the
empirical distribution of the real population. What do we mean by reason-
able? If our aim is to make inferences from a sample about a population
with certain precision and confidence, a reasonable approximation means
that the actual departures of the data from the theoretical assumptions
do not seriously affect the inferences we draw from that data. Fortunately,
most commonly used statistical techniques tend to be fairly robust with
respect to minor discrepancies between data and the theoretical distribu-
tion used to model them, but they are by no means insensitive to major
departures from the assumptions. For example, if daily wages display a
skewed distribution or reveal the presence of significant outliers, infer-
ences based on the assumption of normality can be seriously misleading.
Hence, in order to make valid inferences from a sample about its popu-
lation it is necessary that the assumptions we make in modelling the
population are reasonably valid in practice. But this raises a thorny issue.
Do we not rely in part on the data to verify whether our assumptions are
valid in practice? If so, do we not adapt models in the light of hints, clues
or tests derived from the data so as to end up with a better specification?
But <loes this approach not entail the danger of getting caught up in a
circle: using our data to improve a model and, subsequently, using the
model to make inferences from the data? There is, therefore, a potential
tension between the two roles that data can play in analysis: testing models
against data and getting ideas from data to improve our models. As
we shall see, this tension is inherent in the practice of modelling data.
As such, it is a major preoccupation for any practitioner engaged in
data analysis.
To avoid confusion, however, let us make clear from the outset that
model specification inevitably involves abstraction and, hence, is always
theory-inspired. Nobody argues that facts speak for themselves. Theory
is the driving force in the process of modelling. But this raises two inter-
related questions. First, how much guidance do we expect from theory in
model development or creation? Second, at which point, if at all, do data
enter the scene? Both questions underlie the debates on the approaches
to modelling.
What guidance do we expect from theory? Here is an interesting quota-
tion from two economists-turned-anthropologists on this issue:
Theories can stimulate new ideas, facilitate the posing of novel, inter-
esting questions and generally guide fieldwork. However, they can also
overdetermine field research and be a barrier to the development of
Model specification and applied research 29
understanding. Discovering the fine line between theoretical guidance
and theoretical overdetermination - between pre-formulated questions
and pre-formulated answers - is no easy task.
(Gregory and Altman, 1989: 20)
Hence, in this view, good theoretical groundwork can open up oppor-
tunities in developing new ideas but it can also act as a barrier. Or, as
the anthropologist Malinowski put it more vividly, 'preconceived ideas are
pernicious in any scientific work, but foreshadowed problems are the
main endowment of a scientific thinker' (cited in ibid.: 11). This is a nice
way of stating a key problem that a researcher typically confronts in empir-
ical work, but it still begs many questions when discussing approaches to
modelling. Few methodologists would disagree with Gregory and Altman
that theoretical guidance is a good thing, while theoretical overdetermi-
nation can act as a hindrance. Nobody prides themselves that they engage
in research equipped with preconceived ideas. The problem remains as
to what we mean by appropriate theoretical guidance. How much is too
much? Indeed, what sorne consider as well-structured guidance, others
may dismiss as gross overdetermination. So, when does theoretical guid-
ance turn into overdetermination?
To dig deeper into this question it is necessary to take a closer look at
what we mean by doing theoretical groundwork for research. Sound
research requires an applied worker to be familiar with the theoretical
debates surrounding his or her research question. Theories differ because
they derive from competing schools of thought or because they present
different positions within a given paradigm. Whichever is the case, solid
research requires that a researcher comes to grips with rival theories which
address his or her research question. The question now arises as to the
role data play in choosing among competing rival explanations. Should a
researcher be committed to a particular hypothesis, cast it in a well-
specified model, and subsequently test this model in isolation against the
data? Or should a researcher approach the data in a more open-ended
fashion, allowing the data to play a role in choosing among competing
rival explanations, so as to end up with a preferred model which appears
most plausible in the light of the evidence? Which of these two approaches
should guide the practice of research? In the next section we look at how
traditional modelling strategies approached this question, while in section
1.4 we turn to modern approaches to modelling data.

1.3 THE ROLE OF DATA IN MODEL SPECIFICATION:


TRADITIONAL MODELLING
Traditional econometric modelling (as well as classical statistics) uncon-
ditionally opted for the first variant: models are to be tested in isolation
against data specifically sampled for that purpose. Consequently, theory
reigns supreme in the business of model specification. This tradition in
30 Econometrics far developing countries
econometrics carne of age in the wake of the Haavelmo-Cowles research
programme initiated in the 1940s, and reached its heyday in the 1960s,
after which it carne under increasing methodological attack in the 1970s
(Morgan, 1990).
In this view, model specification was the exclusive preserve of theoreti-
cal groundwork. Data entered the scene to test whether a model stood up
to scrutiny and to estimate its unknown coefficients. If a model failed the
test, a researcher had to go back to the drawing board, reassess the core
theory of the model or its auxiliary premises, and come up with a new con-
jecture - a new model - to be tested against a fresh set of data. Data could
reject or validate a model and estimate its coefficients, but they should
never be allowed to suggest new or better models. Data analysis, therefore,
was confined to what the philosopher R. W. Miller (1987: 173) called 'a
lonely encounter of hypothesis with evidence'. Any attempt to allow data
to play a role in model specification was dismissed as non-rigorous and
methodologically unsound (Heckman, 1992: 884). This amounted to data
mining, which was the greatest sin any researcher could commit.
So much for the principles laid down by this approach. But where did it
leave an applied researcher? In theory, at least, the tension inherent in the
dual role of data was settled by denying data any role in model specifica-
tion. Applied econometrics, as Mary Morgan (1990: 263) put it, was reduced
to a mere statistical complement of theoretical discourse. But applied
econometricians found it hard to conform with this principle, even when
they agreed with it in theory. Real data always prove to be unruly when a
model is first tested against them. A relation assumed to exist between two
or more variables may turn out to be weak or non-existent; the estimated
coefficients of a model may have the wrong signs or take on implausible
values; or the residual variation of the estimated model may signal that the
assumptions about the error term are by no means satisfied. A researcher
adopting the traditional strategy to modelling should in principle discard
the data set, think again and come up with a new model to be tested against
a fresh sample of data. But data are hard to come by: there is only one set
of national accounts for a particular country over a given period; a house-
hold budget survey is normally done only once every five or ten years.
Researchers who collect their own data are equally aware of the costs and
time involved in sampling a fresh set of data. In practice, therefore, the
option of starting anew when a model fails to perform well is rather
limited. So how did applied workers resolve this problem?
Actual practice involved a process of trial and error in which data
inevitably played a role in model specification. Typically, a researcher
would start by trying out a preferred specification against the available data.
If things went well - that is, the results were in line with theoretical expec-
tations and the estimated relation carne out strong - the process could stop
there. But more commonly things did not go smoothly on the first trial. In
this case, it was common to assume that the core idea of the model was still
Model specification and applied research 31
correct, but further complications needed to be taken into account to bring
out the assumed relation in full view. A researcher would try out a range
of ad hoc modifications to the model: perhaps add another variable, change
the functional form of the equation, or incorporate a more complex speci-
fication of the random componen t. This process is not unlike the way a cook
tries out a given recipe, tastes it and, depending on the taste, decides to
add a few more ingredients to improve its aroma or appearance.
This approach to modelling is still very widespread in applied work
today. You might be inclined to argue that cooks who proceed in this trial
and error fashion often make dishes worth eating. This is undoubtedly
true. Similarly, many data analysts often come up with interesting models
arrived at by ad hoc modifications of an initial simpler version. But this
approach also has its pitfalls. To understand this, remember that tradi-
tional econometric modelling requires a researcher to make a firm
commitment to a particular model specification. The model is subsequently
tested against the data in isolation from possible rival models. Most
commonly the model chosen by a researcher is his or her pet theory. Trial
and error through ad hoc modifications will then result in a model spec-
ification which generally retains the main thrust of the researcher's
preferred theory. But this entails the real danger that applied work merely
boils down to 'an unstinting collection of evidence in support of one's
ruling theoretical position' (Pelta and Pelta, 1978: 283). At no point is
any attention given to the possibility that rival theories or models may
explain the phenomenon in question just as well, or even better, without
perhaps requiring as many ad hoc alterations in model specification as
were necessary to get one's pet theory into shape.
Hence, the real problem with this approach is not that it is unable to
come up with interesting models. The problem is that it concentrates
exclusively on one theoretical explanation in isolation from its possible
rivals. But should empirical analysis just boíl down to a lonely encounter
of a hypothesis with its evidence? Or is genuine analysis comparative in
nature? It is perfectly possible that a particular explanation looks plau-
sible in the light of its evidence, yet another rival explanation may do just
as well, if not better. What matters in applied analysis, therefore, is that
empirical analysis enables us to discriminate between rival explanations.
One philosopher put this point as follows:
No encounter with data is a step towards genuine confirmation unless
the hypothesis does a better job of coping with the data than sorne
natural rival ... What strengthens a hypothesis, here, is a victory that
is, at the same time, a defeat for a plausible rival.
(Miller, 1987: 176)
The underlying principie here is that the strength of an idea, or of a model,
shows itself only when compared with rival ideas or models, and not when
analysed in isolation.
32 Econometrics far developing countries
In conclusion, traditional econometric modelling requires a researcher to
make a firm commitment to a particular model which subsequently is to be
tested against data. In this view, therefore, theoretical guidance means that
model specification is the exclusive preserve of theory. In practice, how-
ever, ad hoc modifications are made to the initial specification in the light
of trial and error testing against a given data set. Data mining, while con-
demned in principle, was (and continues to be) rampant in practice. Models
emerge after an iterative process in which data play a role in deciding on
the final specification. Throughout the process, the model remains insulated
from rival explanations and is fortified by a battery of ad hoc alterations
to fit its evidence better. While this process can undoubtedly produce
challenging and interesting models, it lacks the real essence of genuine
testing of a theory which inevitably involves taking account of competing
theories. Because of this critica! weakness, this approach carne under
increased attack, particularly from the 1970s onwards. The traditional
approach subsequently became known as what Leamer called the 'average
economic regression' or what Hendry labelled 'specific to general model-
ling'. Let us now look at sorne modero modelling strategies.

1.4 THE ROLE OF DATA IN MODEL SPECIFICATION:


MODERN APPROACHES
Modero modelling strategies all share the common characteristic that they
tend to be more data-centred, meaning that they allow data to play a
more prominent role in model specification. While traditional modelling
practices saw model specification as the exclusive domain of theory and,
therefore, outside the reach of data analysis, modero approaches to model-
ling include specification searches, guided by theory, as an integral part
of data analysis. More specifically, a common feature of the more modero
approaches to modelling is that they are far more comparative in nature.
Data are made to play an active part in choosing among rival specifica-
tions. In other respects, however, these approaches differ markedly from
one another and, as such, they do not represent a unified strategy for
modelling data.
lt is not our intention here, however, to provide an exhaustive survey
of modelling strategies in statistics and econometrics nor to compare
different methodologies in any great detail (see, for example, Diaconis,
1985; Granger, 1990; Kennedy, 1992: 73-90; 278-89). Instead, in this
section we shall briefty discuss sorne leading ideas and principles of three
main variants of modero approaches to data analysis: (a) general to
specific modelling based on model selection through hypothesis testing;
(b) exploratory data analysis as a tool to detect meaningful patteros in
the data; and (c) fragility (or sensitivity) analysis which checks how sensi-
tive particular results of model estimation and testing are with respect to
neighbouring (competing) model specifications.
Model specification and applied research 33
General to specific modelling
The principle of testing downwards rather than adjusting upwards seeks
to make the problem of model selection an integral part of statistical
inference. It arose in opposition to the established practice in traditional
econometric modelling which, as we discussed in the previous section,
proceeded by making ad hoc alterations whenever needed to an initial
preferred (and simpler) version of a model. The problem with adjusting
upwards is that subsequent statistical inferences ( estimation and hypoth-
esis testing) lose precision because the model has been adapted to fit the
data against which testing and estimation takes place. General to specific
modelling, in contrast, proceeds by testing downwards inasmuch as it
requires a researcher to start with a broadly specified overarching model
which contains within it several feasible specifications that depict rival
explanations of the phenomenon in question. It is akin to fitting a loose
dress to a person - a dress which could fit several candidates - and, subse-
quently, tailoring it down to size. Hence, the process involves two steps.
First, a researcher formulates a model which encompasses rival explana-
tions deemed relevant in the light of theoretical research. The task is to
make sure that the initial broader model is itself an adequate specification
of the data-generating process. Hence, before testing downwards can start,
it is necessary to subject the general model to a battery of misspecification
tests which seek to verify that the noise component of the general model
no longer conveys hidden messages. To prevent this happening, researchers
are advised to make the initial general model sufficiently broad so as to
avoid leaving out relevant factors. The obvious implication is that the ini-
tial specification will inevitably carry a lot of extra baggage.
Second, the researcher then attempts to simplify the general model by
imposing restrictions on it, the validity of which can be formally tested.
In this way, a researcher hopes to arrive at a simpler model which
is acceptable in the light of the empirical evidence. Hypothesis testing
against data plays an active role in model selection which involves a choice
among rival models. Testing downwards, therefore, implies general to
specific modelling through vigorous sequential hypothesis testing which
enables a researcher to weed out rivals which do not stand up to the
empirical evidence and to zoom in on a model which appears plausible
in the light of the data.
Consider a simple example. A researcher postulates a model which
involves a simple relation between Y and X along with a noise element
(the error term). A rival theory, however, suggests that a third variable,
Z, is also important in explaining the variation in Y. Traditional econo-
metric modelling proceeds by first trying out the simple relation between
Y and X. If the results look good, our researcher will stop there. General
to specific modelling, however, requires that the researcher starts with the
broader model - Y is explained by both X and Z along with an error
34 Econometrics far developing countries
term - and subsequently the researcher should test formally whether Z
can be dropped from the equation. Hence, our researcher should formally
test the restriction that the coefficient of Z equals zero against the data.
The simple model is only selected if the data allow this restriction to be
imposed on the more general model.
This approach of general to specific modelling puts great emphasis on
rigorous testing: misspecification testing of an initial general model
followed by further model selection based on sequential testing of various
restrictions imposed on the general model. For this reason, this approach
(which is associated with work carried out by Professor Hendry at the
London School of Economics in the 1970s and 1980s) is often referred to
as the test-test-test approach to modelling because of its insistence on
rigorous testing (see for example, Gilbert, 1990; Spanos, 1990; Charemza
and Deadman, 1992; Kennedy, 1992: 73-90). The approach has been devel-
oped mostly in the context of dynamic models involving time series
analysis, but the general principle of model selection by testing down-
wards is obviously equally applicable to cross-section data.
How <loes this approach address the tension between leaming from data
and testing ideas against data? To resolve this tension, general to specific
modelling integrates model selection within the realm of statistical
inference. To leam from data which model to choose, it is necessary
to select between models through formal hypothesis testing. As we shall
show repeatedly in this book, this approach to modelling is a handy and
powerful tool for applied workers. Whenever possible, it is always prefer-
able to test downwards rather than adjust upwards so as to check the
validity of the assumptions underlying a particular model.
But as is the case with all methodological approaches, the method of
general to specific modelling also has its weaknesses, apart from its consid-
erable strengths as a guide to data analysis. Two problems are worth
mentioning in this context.
First, the particular path taken in testing downwards from a general
specification to a specific model may determine the final outcome. In other
words, two researchers using the same general model with the same data
will not necessary arrive at the same restricted version of the model. A
given set of data may well admit quite different theoretical explanations.
This in itself should not surprise us; empirical evidence does not always
allow us to discriminate clearly between rival explanations. Part of the
problem is that, by its nature, a general model carries a lot of extra
baggage. With non-experimental data, it is quite common that the different
variables in an equation end up blurring each other messages because
they tend to overlap in various ways. As Leamer (1978) argued, much of
the testing downwards exercises undertaken in applied work have a lot
to do with making sense of blurred messages and involve interpretative
or simplification searches rather than clear-cut hypothesis testing between
well-defined rival explanations. That is, restrictions are imposed on the
Model specification and applied research 35
general model to ease the economic interpretation of the results (because
the data do not allow for greater complexity). In such circumstances
testing downwards may well be more a matter of convenience - the hope
of arriving at a restricted version which makes good sense - than of a
conscious choice between rival explanations.
Second, the method of model selection by testing downwards starts with
the proposition that, as Heckman (1992: 883) put it, 'a wide class of models
can be, or has been, enumerated in advance of looking at the data and
that empirical work consists in picking one element in a fixed set'. In other
words, the theoretical groundwork prior to data analysis is supposed to
come up with a comprehensive account of all rival theories deemed rele-
vant to the research question and to cast them within the confines of a
testable overarching model which allows data analysis to play its part in
the process of selection between this fixed set of rivals. But, with
Heckman, we can argue that 'more often, empirical work suggests rich
new classes of models that could not have been anticipated befare the
data are analysed' (ibid.). The point is that data analysis can enrich theory,
not just pick a rival explanation among a predetermined list.
This second point is highly relevant for development economists. If our
interest is to estímate a demand function for food, for example, we can rely
on demand theory and a wide range of empirical examples to guide our
work. In this context it is preferable to make sure that we start with a
general specification which includes all variables deemed relevant and, sub-
sequently, proceed by testing downwards. In this case our aim is well
defined (estimating a demand function) and economic theory provides us
with strong guidance. But what if we embark on a study of the effects of
structural adjustment on informal sector manufacturing in a developing
country? In this case we may find that our initial theoretical preparation
does not give us such firm handles that we can integrate into a model which
embraces all relevant rival explanations. Theory will provide guidance, but
more in the shape of a set of vague ideas - avenues to be explored - rather
than a set of fully fiedged rival hypotheses. Research of this nature will
almost certainly involve exploratory empirical work of a type which
falls outside the reach of formal statistical inference. This brings us to
exploratory data analysis (EDA) as a second approach to modelling data.

Exploratory data analysis (EDA)


The set of techniques initiated by Tukey (1977) and Mosteller and Tukey
(1977) rapidly evolved into a novel approach to data analysis. EDA
puts the emphasis squarely on learning from data so as to arrive at an
explanation which appears plausible in the light of the evidence. At first
sight it may appear that this approach simply advocates facts speaking
for themselves and, hence, that it is sufficient to scrutinise data to distil
an explanation out of empirical evidence. This is not the case. Data by
36 Econometrics far developing countries
themselves do not tell you anything unless you engage with data in a
dialogue which is theory-inspired. It is necessary for you to fire questions
at the data so as to get hints and clues from them. In exploratory data
analysis, this process of questioning the data has two distinct features.
First, EDA insists that data should be approached from different angles
by analysing them in several different ways through an interactive process
informed by theoretical refiection (Hamilton, 1992: viii). The point is to
try out a particular avenue, refiect on it, perhaps follow up any clues or
hints, and plan what to do next. It involves, therefore, a trial and error
process inspired by theoretical preparation and subsequent refiection. The
process will obviously be richer, the more a researcher incorporates the
perspectives of rival theories in the analysis of data so as to weed out
theories which appear implausible in the light of the evidence and to
follow up more promising avenues.
Second, in pursuing a given question, EDA proceeds by fitting patterns
readily detectable in the data so as to be able to focus on the remaining
residuals obtained by removing the fit from the data. Hence,
DATA= FIT + RESIDUALS

where the FIT is a preliminary model employed to remove key patterns


from the data so as to be able to look more carefully at the residual
variation. In EDA, a researcher <loes not assume that the residuals no
longer contain meaningful messages and, hence, they cannot be put aside
as pure noise.
The latter point is important. In traditional econometric analysis, resid-
uals played only a relatively minar role. At most, a few tests were routinely
performed with residuals to check whether it was necessary to make
specific ad hoc modifications to the original specification. In this view,
residual analysis allowed a researcher to detect the symptoms of a few
known specification illnesses which could readily be cured by applying
appropriate corrections. Modern approaches to modelling do not operate
in this way. If residuals behave contrary to model assumptions, modern
approaches to data analysis will see this as a general sign of model misspec-
ification. This is the reason why general to specific modelling, based on
the principle of test-test-test, always starts with a battery of misspecifica-
tion tests to check whether the general model is data admissible. EDA,
in contrast, uses residuals to look for further meaningful patterns in the
data so as to come up with novel discoveries as to which other factors
should also be taken into account. Hence, in EDA, the basic premise is
that it is necessary first to remove patterns which are obviously present
in the data so as to bring out more hidden factors at play. To put it simply,
while traditional modelling often finds the residual variation a nuisance
(particularly when it fails to behave as pure noise ), exploratory data
analysts treat residuals of earlier fits as a rich source from which to extract
clues and hints as to further avenues of inquiry. In this way they hope to
Model specification and applied research 37
be able to come up with new models which could not conceived befare
the data are analysed.
To find the unexpected, it is necessary to look for it. This is the reason
why EDA, unlike traditional modelling approaches, makes extensive use
of analytical graphics along with numerical summaries. As Kennedy (1992:
284) put it:
EDA, exploratory data analysis, is an approach to statistics which
emphasizes that a researcher should begin his or her analysis by looking
at the data, on grounds that the more familiar one is with one's data
the more effective they can be used to develop, test, and refine theory.
Econometricians are often accused of never actually looking at their
data. Exploratory data analysts believe in the inter-ocular trauma test:
keep looking at the data until the answer hits you between the eyes!
EDA, therefore, allows us to get clues and hints from data which, after
further theoretical reflections, may lead us to new ideas and models. In
terms of the inherent tension between learning from data and testing
against data, EDA is explicitly concerned with the former. This is its
strength, but also its weakness. In situations where our knowledge is as
yet vague and no firm hypotheses exist, EDA can be of great help to
dig deeper into a problem. At times, due to lack of data availability, this
may imply that we cannot do more than arrive at a model which seems
plausible in the light of the evidence without being able to test it inde-
pendently against a fresh set of data. But at times it is possible that ideas
obtained through data exploration in one part of the research lead us to
interesting follow-ups which can then be formally tested against the data.
In actual research exploratory and confirmatory data analysis interact in
ways which are neither purely rigorous hypothesis testing nor mere post-
data model construction.

Fragility or sensitivity analysis


How sensitive are the key results of model estimation to minar changes
in model specification? This is a question which guides a third approach
to modelling rooted in fragility or sensitivity analysis. This approach is
associated with the work of Leamer (1978, 1983); a similar notion can also
be found in Mosteller and Tukey (1977) in their discussion on all-subset
regression and on the woes of regression coefficients. The basic principie,
as Leamer (1983; Granger, 1990) put it, is that an inference is not believ-
able if it is fragile, i.e. if it can be reversed by minar changes in assump-
tions. The point is that often researchers are mainly interested in the value
or the sign of one coefficient in a model. For example, a researcher may
wish to know whether the demand for a particular commodity is price elas-
tic or not. Alternatively, the interest of a researcher may be to investigate
whether public investment crowds out private investment. Or you may
38 Econometrics far developing countries
wish to test Griffin's famous hypothesis that fareign aid displaces domestic
saving. In all such cases, the attention of the analyst centres on the value
or sign of the coefficient of what Leamer referred to as the facus variable
in the model. But the models actually employed to obtain these estimates
will generally include more variables, the coefficients of which are of lesser
immediate concern. If, however, minar changes in model specification lead
to majar changes in the coefficient of the facus variable, we have good rea-
son to doubt the conclusions drawn from any such particular specification
which proves to be highly fragile.
Leamer's methodology is Bayesian in approach and hence, by its
very nature, more comparative in scope than classical statistical inference.
Leamer's particular brand of fragility analysis, or what he called extreme
bounds analysis, is highly complex. The 'extreme bounds' refer to the upper
and lower boundaries of the range within which a coefficient varies with
respect to alternative model specifications. Leamer's specific approach is
well beyond the scope of this book, but we shall nevertheless retain the
general idea of fragility analysis: the necessity to investigate the bounds
within which a coefficient varies as a result of minar changes in model
specification.
The basic idea, therefare, is that key results of a model should be
subjected to sensitivity analysis with respect to a neighbouring range of
alternative specifications which differ from one another by the inclusion or
exclusion of a number of doubtful variables (variables which the researcher
is not fully confident should be included or not in a specification), apart
from key variables (one or more of which can be facus variables) which
feature in all specifications. If the inferences from the coefficients of facus
variables prove to be highly fragile with respect to the inclusion or exclu-
sion of doubtful variables, not much confidence can be placed in the
conclusions derived from such inferences. This approach to modelling,
therefare, seeks to assess how fragile the inferences we make are with
respect to minar changes in model specification. To do this, we seek to find
the bounds in the variation of one or more key coefficients across a range
of alternative neighbouring specifications so as to judge the robustness of
our results to minar changes in model specification.

Summary
In this section we reviewed three distinct sets of ideas on the role of data
in model specification. These three methodologies share in common that
they assign an active role far data in model choice, development or selec-
tion. Obviously, as shown above, each approach has its distinctive ftavour
which corresponds to a particular methodological outlook. But we would
also argue that each approach has its own strengths and weaknesses which
differ depending on the particular context of research. If, far example,
your particular interest is to show that public investment crowds out
Model specification and applied research 39
private investment, it is not very sensible to settle on one specification
which happens to produce the required negative coefficient if minar alter-
ations to this preferred specification render this result insignificant or even
reverse its sign (we shall see an example of this in Chapter 6). If you are
dealing with a problem of model choice in which economic theory provides
you with forceful handles that allow you to nest rival models inside a
more general specification, hypothesis testing in the context of general to
specific modelling seems to be a logical choice of modelling strategy. But
if, in contrast, your research question is still rather vague and theory can
do no more than indicate plausible avenues of inquiry, a researcher may
well have to rely on extensive data exploration to arrive at a firmer hypoth-
esis. In actual practice, from the start of a piece of applied analysis to
its conclusion you may well find yourself drawing on a combination of
the three approaches.
Hence, while each of these approaches is rooted in distinctive method-
ological outlooks, it seems fair to say that each approach will prove its
strengths or reveal its weaknesses, depending also on the specific context
of the research. Therefore, in this book, we do not seek to rally your
support in favour of one of these approaches, but instead our aim is to
draw upon each of these methodologies and show their usefulness in
different research contexts so as to enhance your own ability to employ
data actively in the process of model specification. Indeed, the three
approaches are complements rather than substitutes, all rooted in the same
basic philosophy that data have a role to play in model specification.

1.5 THE TIME DIMENSION IN DATA


The data we use in development studies always have a time dimension.
Time enters in a double sense depending on (a) how we measure our vari-
ables, and (b) whether our interest is to study evolution over time or to
make comparisons at a particular point or period in time.
As to measurement, it is useful to distinguish clearly between stock and
fiow variables. A stock is measured at a particular point in time; a fiow
is measured for a particular period (say, a month or a year) in time. The
consumer price index, for example, is a stock variable since it measures
the level of prices at a point in time. The rate of inflation, in contrast, is
a flow variable defined over a period in time. Similarly, population is a
stock variable, while births, deaths and migration are flow variables. These
examples also show that changes in stock variables are themselves fiow
variables. Investment, for example, is a fiow variable which adds to the
stock of capital.
Time series depict evolution over time; cross-section data make compar-
isons at a particular point or period in time. National income accounts,
trade data, industrial production statistics and price series are examples
of time-series data. A household budget survey, a population survey, an
40 Econometrics for developing countries
industrial survey, employment and labour market surveys are typical
examples of published cross-section data. Both types of data are commonly
used in development research, although modern-day econometrics tends
to give greater prominence to the analysis of time series data.
Cross-section data generally consist of surveys and, hence, involve
random sampling from a wider population. Regression theory is based on
the assumption that the sample is drawn randomly from the population,
and so survey data normally conform with this requirement. In contrast,
however, time-series data seldom do, and therefore they pose special
problems.
Indeed, a time series can best be seen as one observation of a partic-
ular historical process over a number of years, rather than as a collection
of independent observations of unconnected occurrences. This year's price
of a commodity bears sorne relation to last year's price. GDP growth,
stagnation or decline takes place in discernible trends, reversals or cycles.
Time-series data, therefore, reflectan inherent momentum which can often
continue over long periods of time. This is the reason why conventional
time-series analysis (Box and Jenkins, 1970) sought to forecast the future
evolution of a variable solely in terms of its own past. In sum, time series
seldom behave as a set of data randomly sampled without any connec-
tion between successive observations. Instead, they reflect the underlying
history of an economy or society.
But this leads us to a second problem inherent in working with time
series: the danger of interpreting spurious correlations as evidence for the
existence of causal connections between two or more variables. What is
a spurious correlation? At its simplest it may mean that two or more vari-
ables are causally unrelated but, by chance, a correlation exists between
them. But in time-series analysis we are likely to observe spurious
correlations for another reason. Consider a simple example. Two persons,
unconnected in any way, walk side by side along the same street in the
direction of a bus stop to catch the same bus. We observe their move-
ments but are unaware of their respective purposes. We might be tempted
to infer that they belong together. This is a spurious correlation. In fact,
each intends to catch the bus, never mind what the other does, and neither
inftuences the behaviour of the other. Yet they walk together at the same
time in the same direction. Time series often 'walk together' in this way:
both may be following a deterministic trend or what is called a random
walk. Hence the fact that many socioeconomic variables seem to move
together should not lead us to infer that such variables are necessarily
causally linked in a direct manner. The history of econometric practice,
however, has shown that all too often such causal inferences are made.
But all is not bad news as far as working with time series is concerned.
Time series also offer opportunities in analysis precisely because of their
time dimension. They allow us to investigate the structure of determina-
tion between variables over time by linking the past history of sorne
Model specification and applied research 41
variables to the future evolution of others. For example, we could inves-
tigate whether, in a cash-crop-producing economy constrained by foreign
exchange availability, income earned from last year's sale of export crops
limits expenditures on imports in the current year. Put differently, time
series allow us to specify equations which include lagged variables so as
to capture dynamic interactions between variables over time. Furthermore,
the inclusion of lags in the dependent variable allows us to distinguish
between long-run equilibrium relations and short-run dynamic disequi-
librium behaviour. This explains why time series play such a prominent
role in econometrics.
Recent developments in econometric theory and practice have been
particularly concerned with these special challenges posed by time series
in applied work. The danger of spurious correlations has led to more
attention being given to modelling trends (particularly, stochastic trends)
in time series. In fact, as we shall see, many economic time series tend
to behave as random walks. This has led to the development of cointe-
gration analysis (Granger and Newbold, 1974, 1977) which seeks to
investigate long-run relations between economic variables without falling
prey to spurious correlations. Furthermore, error-correction models (for
a recent and comprehensive reference, see, for example, Banerjee et al.,
1993) seek to incorporate short-run disequilibrium behaviour along with
long-run tendencies in econometric modelling.
A word of warning is, however, in arder with respect to empirical
analysis with time series. In econometric models, time generally enters as
a subscript of variables in an equation. That is, time enters as logical time,
not chronological or historical time. A useful way to see how time is dealt
with in econometric studies is to consider an analogy with experiments in
physics. Suppose we heat a container with a fixed volume of water and
graph the rise in the temperature of the water against time (measured,
say, in seconds). We can repeat this experiment severa! times under iden-
tical conditions to produce the same graph over and over again. Time t
in the graph refers to logical time, not historical time. Development
research, however, is a historical science. Time is more than a mere
subscript in an equation. Economies and societies evolve and transform
over time. Events unfold over time and produce irreversible change. For
example, South Africa in the 1990s is qualitatively different from what
it was in the 1960s. We can analyse history but we cannot replay it.
Econometric models often aim to come to grips with structural interac-
tions between variables which characterise the basic momentum of an
economy over a prolonged period in time. In so doing, we abstract from
the more historically specific events which marked this particular period,
but obviously we cannot abstract from the broader historical context
to which the model refers. The point of it all is to model sorne key
features which capture the interactions between variables inherent in the
momentum of the economy in the period concerned. Note, however, that
42 Econometrics far developing countries
data exploration of time series can help to bring out specific historical
features which matter in modelling the dynamic character of a particular
economy in a particular period.

1.6 SUMMARY OF MAIN POINTS


1 Modelling data is not just a question of model estimation and hypoth-
esis testing; it also involves model specification. Traditional textbooks
in statistics and econometrics tended to emphasise the former, but
applied researchers are often more concerned with the task of finding
an appropriate model which best answers the research question.
2 A statistical or econometric model seeks to come to grips with rela-
tions between variables in non-deterministic situations in which
regularity in data (an average or an average relation) goes hand in
hand with considerable residual random variations. A model specifi-
cation, therefore, has both a structural component (sound) as well as
a residual error component (noise), both of which together seek to
capture the variation in the data. Modelling data, therefore, requires
that we rely both on substantive (development economics) and statis-
tical (probability) theory.
3 The validity of statistical inferences (estimation and hypothesis testing)
depends on whether the assumptions behind the models we use to
analyse the data are reasonably valid in practice. Befare making any
inferences, therefore, it is necessary to check whether the assumptions
of our model are reasonable, given the problem at hand.
4 There is a potential tension between the dual roles data can play in
analysis: (a) testing ideas against data within the confines of a given
model, and (b) getting ideas from data to develop or improve a
model or to choose between rival models. Modelling strategies differ,
depending on the way they handle this tension.
5 Traditional modelling approaches sought to resolve this tension by
denying that data had a role to play in model specification. Theory
reigned supreme in the business of model specification. Data entered
to test a model in isolation from its rivals or to estímate its coeffi-
cients. In practice, however, applied work generally involved making
ad hoc modifications to a model in the light of the initial results of
model estimation and testing. Data mining, while condemned in prin-
ciple, turned out to be rampant in practice.
6 Modern approaches to modelling give data a greater role to play in
model specification: (a) general to specific modelling allows data to
choose between rival models through formal hypothesis testing. Model
selection takes place by testing downwards within a general model
which allows far different interpretations depending on the nature of
the restrictions imposed on it; (b) exploratory data analysis seeks to
obtain hints and clues from data by looking carefully at the (residual)
Model specification and applied research 43
patterns within the data. Its main object is to learn from data, rather
than to test ideas against data; (c) fragility analysis investigates how
sensitive the results of model estimation and hypothesis testing are
with respect to minar changes in model specification. The basic idea
behind this approach to modelling is that not much confidence can be
placed in fragile inferences which can be reversed by minar changes
in the model.
7 In data analysis it is useful to distinguish between stock and flow vari-
ables. The former are measured at a point in time and the latter are
measured over a period in time. Stock and flow variables can feature
in both cross-section and time-series analysis. Cross-section data make
comparisons at a given point or in a given period in time, while time-
series data depict evolution over time. The latter are specifically suited
to analysing the dynamic interplay between variables but pose the
problem that they are prone to spurious correlations.
2 Modelling an average

2.1 INTRODUCTION
This chapter and Chapter 3 <leal with the problem of modelling a simple
average of a single variable. But why bother with univariate analysis if,
in development research, our main interest is to study empirical relations
between two or more variables? Why not jump straight to regression
analysis? We can think of three reasons why it is best to start with uni-
variate analysis.
First, we should always be aware that specific features we may come
across in univariate analysis, such as the presence of an outlier or of pro-
nounced skewness in the distribution of a variable, invariably have multi-
variate implications. Unexpected or puzzling results in regression analysis
can often only be properly understood if we look at the distributions of
its variables. Failure to do this often leads to nonsense regressions. For
this reason, it is best to proceed from the ground up: study each variable in
turn befare embarking on investigating relations between them (Hamilton,
1992: 1-2). Starting in this way also gives you an excellent opportunity to
become familiar with basic techniques of EDA (exploratory data analysis)
which teach you how to look carefully at a batch of data. Second, residuals
play a key role in the process of modelling data with regression analysis, par-
ticularly in the context of modern modelling strategies. To verify whether
the assumptions of the models we use are valid in practice, it is important
to look carefully for hidden messages in the residuals. To do this, we treat
the residuals of a regression as an observed variable in its own right.
Univariate analysis helps us to look for patterns within residuals or to test
the distributional assumptions we make about the random error term in a
regression. Finally, regression analysis involves averaging of a complex
nature. In empirical analysis, when we say that Y is a function of X, we
mean to say that the average value of Y is a function of X (Goldberger, 1991:
5). In other words, in regression analysis we <leal with conditional means of
Y for given values of X. Consequently, common errors we make when
dealing with a simple average often crop up again in more complex forms
when we subsequently move from univariate to multivariate analysis.
Modelling an average 45
Data analysis embraces both the problem of finding an appropriate
model (model specification), on the one hand, and model estimation and
testing, on the other. This chapter only deals with the latter aspect: esti-
mation and hypothesis testing within the confines of a model which we
assume to be correct. It further assumes that a univariate sample is drawn
from a normal distribution. Why do we make this assumption? One reason
could be that most data we encounter in practice are approximately
normal. Hence, if normality is the rule, it makes good sense to start with
this assumption. Unfortunately, while in sorne sciences data often behave
in this way, most social or economic data are not (approximately) normally
distributed. As we shall see, it is hard to find examples of social or
economic data which display the typical bell-shaped, normal distribution.
More often than not, social and economic data are skewed. Another
reason could be that the normal distribution is ideal for obtaining mean-
ingful averages and, hence, serves as a useful example for the problem of
averaging. This is indeed the case and explains why we take the normality
assumption as a point of our departure. A final reason is that it is often
possible to find an appropriate mathematical transformation which elim-
inates skewness in the distribution of a variable, and makes the normality
assumption acceptable.
In section 2.2, we show intuitively that normality in data renders it
easier to make sense of averages. Section 2.3 then reviews the assump-
tions of the classical model for estimating the mean of a univariate
distribution. Section 2.4 shows that, given these assumptions, the arith-
metic mean of the sample data is the best, linear, unbiased estimator of
the population mean. Subsequently, section 2.5 introduces the principie
of maximum likelihood and shows that the sample mean is also a
maximum likelihood estimator if the population distribution is normal.
Section 2.6 then deals with estimation and hypothesis testing with respect
to the population mean. Finally, section 2.7 summarises the main points
of this chapter. What to do if the normality assumption is not valid in
practice will be dealt with in Chapter 3.

2.2 KINDS OF AVERAGES


In data analysis, we use theoretical abstractions - probability distributions
- to analyse real data. Model specification, therefore, should capture the
main features of the data. An average is one such feature: it gives us the
location parameter of a distribution (Rosenberger and Gasko, 1983: 297).
To model an average, it is useful to refiect on the kinds of averages we
use when dealing with real data, and what each of these averages tell us
about the data. This type of refiection should help us to gain insights into
the problem of modelling an average.
In descriptive analysis, three kinds of averages are frequently used: the
mean, the median and the mode. To obtain the mean of a sample, we
46 Econometrics far developing countries
sum the data and divide the result by the number of observations. The
median is the middle value of the ordered list of the data: half of the
observations lie below the median and half above it. The mode is the
most frequently observed value of the data. In a theoretical distribution,
the mean and median give us respectively the centre of gravity (i.e. the
balance point: the point at which a distribution would balance if it were
made of a solid substance) and the centre of probability (middle value)
of a distribution, while the mode is its peak. If a distribution has several
peaks (not necessarily of equal height) we say that it is multimodal.
An average, taken alone, tells us very little about the data. To interpret
an average we need to have a good idea about the shape of the empirical
distribution of the data which shows the pattern of variability in the data.
To look at shape, we shall make use of the familiar histogram. Generally,
the questions we are interested in are the following: (a) are the data uni-
modal, or multimodal? (b) are the data symmetrically distributed around
the average, orare they skewed? (c) does the empirical distribution have
a main body together with fat, thin or no tails? and (d) if unimodal and
symmetric, does the distribution display a typical bell-shaped curve?
These questions help us to search for key features of an empirical distri-
bution and allow us to assess the relative usefulness of different kinds of
averages. If, say, a distribution turns out to be bimodal, a median or a
mean will be of little use. In this case, it is preferable to look at each
mode in turn. More often than not, bimodal (or multimodal) empirical
distributions give us a clear signal that we have lumped together two (or
more) sets of data which should have been kept separate. For example,
if women are paid less than men, income data for a given level of educa-
tion or skill may turn out to be bimodal. In general, whenever a
distribution shows clear multimodal patterns, sit back, think hard, and try
to find the factor(s) which may account for the multiple modes. In this
way, the data can often be split up in distinct sets of unimodal data (which
are always far easier to analyse ).
But even if our data are (roughly) unimodal, shape continues to matter
when looking at the practica! relevance of different kinds of averages. In
general, symmetry in the data makes life easier for a data analyst. If, fur-
thermore, the data turn out to be bell-shaped, the notion of an average
becomes even more meaningful. To see this, take a look at Figure 2.1. The
top panel depicts the daily demand (in terms of number of workers) for
casual (manual) labour in Maputo harbour, from March 1980 to June 1981
(a total of 485 observations); the bottom panel gives the distribution of
actual recruitment of these workers for the same period. The data for this
example are in the data file MAPUTO. Both distributions are reasonably
symmetric and bell-shaped (with the data on recruitment showing a slight
skew to the left). For this reason, we superimposed a normal distribution
(with the same mean and standard deviation as the data) on each of
the histograms. The vertical lines depict the means of both distributions.
Modelling an average 47
0.25
0.2
e:
o 0.15
·u
~ 0.1
LL

0.05
o
o 100 200 300 400 500 600 700 800 900 1000 1100
Demand for labour: day shift

0.25
0.2
e:
o 0.15
ü
~ 0.1
LL

0.05
o
o 100 200 300 400 500 600 700 800 900 1000 1100
Recruitment of labour: day shift
siaia. . .
Figure 2.1 The demand for and recruitment of casual labour, Maputo harbour

Table 2.1 lists the means, medians and modes, along with the standard
deviations, for both sets of data. Since the mode is the midpoint of the
group with the highest frequency (fraction) of the data, computation of its
location is sensitive to the number of groupings used to construct the
histogram. As we can see, in each case the mean, median and mode are
virtually equal to each other. Hence, for descriptive purposes, it <loes not
matter much which one we use. They all tell the same story.
In sum, this type of bell-shaped distribution has the following charac-
teristics:
1 The average (mean, median or mode) is unambiguously located in the
centre of the distribution. Symmetry assures that the two halves left
and right of the average are mirror images.
2 The greater the distance from the average, the lower the frequency:
the mass of the distribution is concentrated in the neighbourhood of
the average.
3 The smaller the variance, the more representative the average becomes
for the data as a whole.
Table 2.1 Averaging the demand for and recruitment of labour at Maputo
harbour
Mean Median Mode Standard deviation
Demand 574 571 Similar 155
Recruitment 500 503 Similar 109
48 Econometrics far developing countries
In our example, average recruitment is considerably lower than the
average demand for labour. Furthermore, the variation in recruitment
clearly does not match the much greater variation in the demand for
labour. In this case, therefore, the supply of labour appears to have been
insufficient and relatively inflexible with respect to the larger variations
in demand. Both examples show us that the average of a bell-shaped distri-
bution is easy to interpret. This ease of interpretation is due to the
essential symmetry of the data.
Note, however, that symmetrically distributed data are not always bell-
shaped. For example, the distribution of rounding errors (say, when we
round up aggregate data to the nearest million) has a typical rectangular
shape - a distribution with a body but no tails. We can best describe this
distribution by its range (the difference between the two extremes) since
its average (mean, median or mode) is nothing but the middle value of
this range.
Now, take a look at Figure 2.2 which depicts a skewed empirical distri-
bution. It is the distribution of weekly overtime payments for casual labour
on the day-shift in Maputo harbour in the period from March 1980 to
June 1981. During this period, wage rates were constant, but obviously
weekly earnings will differ due to variations in recruitment and in access
to overtime work. The data were obtained by taking a random sample
of the weekly earnings of 45 workers over 13 weeks randomly selected
within this period, and subsequently selecting those observations (368 in
total) which pertained to the day-shift (as distinct from the night-shift).
The vertical lines in the graph show, from left to right, the locations
of the mode, median and mean of the distribution. Table 2.2 lists their

0.3

0.2
e
o
u¡:i
LL

0.1

o
o 500 1000 1500 2000 2500 3000 3500
Overtime payments, day shift
s1a1a'"
Figure 2.2 Weekly overtime payments
Modelling an average 49
Table 2.2 Averaging overtime payments
Mean Median Mode Standard deviation
Overtime payments 674 525 ±80 629

numerical values, along with the value of the standard deviation. As can
be seen from Figure 2.2, these three kinds of averages do not tell the
same story. In this case, therefore, the usefulness of an average is much
more ambiguous.
In fact, the mean, median and, particularly, the mode are far apart.
Each of these 'averages' tells a different story. The mode is the peak of
the distribution, the median its middle value, and the mean its balance
point. This example shows that, for a unimodal distribution which is
skewed to the right (i.e. its tail is on the right), the mode will be smaller
than the median which, in turn, will be smaller than the mean. Conversely,
the mode of a unimodal distribution skewed to the left will be greater
than its median which, in turn, is greater than the mean. The lack of
symmetry, therefore, results in no clear centre of the distribution. In fact,
in this example, the most distinctive feature of the distribution is its virtual
exponential decline from left to right, ending up in a long tail.
In sum, when an empirical distribution is unimodal, symmetric and
(preferably) bell-shaped, the concept of an average is fairly straightfor-
ward. The peak of the distribution, its middle value and its balance point
all coincide. Symmetry ensures that both halves left and right of the
average are mirror images. The bell-shaped distribution implies that the
frequency declines as the distance from the average increases. By contrast,
the average of a unimodal skewed distribution is far more ambiguous: the
location of the mean, the median and the mode depends on the distrib-
ution of the data between its peak and its tail.
Let us now return to the question of modelling data. The lesson we can
learn from these examples with real data is that modelling an average
invariably requires us to make assumptions about the shape of the distri-
bution. In this chapter we shall assume that the population distribution is
symmetrical and bell-shaped. More precisely, we shall assume that the
relevant model is the normal distribution. In this case, the population
mean (the first uncentred moment of the distribution), the median and
the mode all coincide in one unambiguous 'average'. Furthermore, the
normal distribution is characterised by its thin tails. In fact, there is only
a 5 per cent chance that an observation drawn from a normal distribu-
tion is more than 2 standard deviations away from its mean; the probability
of encountering an observation which is more than 3 standard deviations
distant from the mean is as low as 0.3 per cent. Finally, as we shall see,
if data are distributed (approximately) normally, the mean and the stan-
dard deviation tell us all we need to know about the data.
50 Econometrics far developing countries
Exercise 2.1
Using the data file SOCECON (with world socioeconomic data for 1990)
on the diskette, make histograms and compute means, medians and modes
for the following variables:
1 GNP (gross national product) per capita;
2 HDI (human development index);
3 FERT (fertility rate);
4 LEXPM and LEXPF (male and female life expectancy);
5 POPGRWTH (population growth rate ).
In each case, discuss the different averages in the light of the shape of
the empirical distribution. Would you say that any of the distributions is
reasonably symmetrical and bell-shaped? (If you do not know how to
compute a median, jump ahead to Box 3.1 in Chapter 3.)

2.3 THE ASSUMPTIONS OF THE MODEL


A statistical model is an analytical construct which helps us to come to
grips with non-deterministic situations in which regularity in the data can
only be understood in a context of considerable random ftuctuations.
Perhaps the simplest statistical model is one where its systematic compo-
nent merely states that the variable fluctuates around a constant
population mean, µ:
Y¡ =µ + E; i = 1 ... n (2.1)
where E; is a random variable which depicts the random fluctuations of the
data around its constant mean. In statistical language, this random variable
is referred to as the error term or the disturbance term of the model. If we
intend to use this specification in practice, the first task we have to confront
is to check whether the assumption of a constant mean is reasonable.
For example, the assumption of a constant mean <loes not seem far-
fetched in the case of the day-to-day fluctuations in the demand for labour
in Maputo harbour, at least in the period specified. But take a look at
Figure 2.3, which plots real manufacturing GDP of Tanzania against time
over the period 1964-73. It is obvious that it would make little sense to
average manufacturing output over that period. The reason is that the
variable in question clearly grows over time and, hence, the assumption
of a constant mean for all observations is simply not valid. Consequently,
the systematic component in equation (2.1) is clearly misspecified in this
case. A more appropriate specification would be:
(2.2)
where t = 1964 . . . 1973. In fact, the systematic component, µ 1, of the
model could conceivably be depicted by a linear trend, as follows:
(2.3)
Modelling an average 51

900
800
ctl
·¡:
ctl
N
700
e:
~ 600
IJ...
o 500
CD
Ol
e: 400

t5 300
.!!!
::::>
e:
ctl
200
~
100
o
64 66 68 70 72 74
Year
Siaia™

Figure 2.3 Real manufacturing GDP, Tanzania

The model now becomes,


Y1 =a + ¡3t + E1 (2.4)
an alternative specification which allows for the mean of Y 1 to vary linearly
with time. But now we are jumping ahead into bivariate analysis, a topic
we shall further explore in Chapter 4. The point we want to make here,
however, is that there are many instances in which we cannot assume that
a constant mean prevails for all observations in the sample. This assump-
tion is questionable in many time-series applications, as it may be with
sorne applications with cross-section data.

Exercise 2.2
Can you think of a few concrete examples with cross-section data where
the assumption of a constant population mean for all observations in the
sample is clearly inappropriate?

In fact, we have come across one example already. If women are paid
less than men for equal levels of education or skills, equation (2.1) would
be inappropriate to model the ftuctuations of income of both men and
women with similar education and skills. In this case, it would be more
correct to apply the model separately to incomes of, respectively, men
and women. Similarly, mortality levels may differ between urban and rural
populations, or among social classes. Averaging across these categories
may well give us misleading results since we assume a single population
when, in fact, several distinct populations should be considered.
52 Econometrics far developing countries
But let us assume that we are dealing with a situation where the assump-
tion of a constant mean is valid in practice. The model as specified in
equation (2.1), however, is still incomplete. The reason is that we need
to specify the stochastic nature of the error term. In classical statistics, at
least three assumptions are made with respect to the behaviour of the
error term:
E(EJ =o (2.5)

(2.6)

E(E;E) = O for all i =fa j (2.7)


The first assumption is obvious: the error term has zero mean. This
ensures that µ is indeed the population mean of the variable Y;, since,
E(YJ = µ + E(EJ (2.8)

Assumption (2.6) states that the error term is homoscedastic. That is,
it has a constant variance. When we say that the error term has a constant
variance, we do not mean to say that all error terms will have the same
size. What it means is that each error is drawn from a population with
the same variance. Since the probability distribution of Y; and E; are iden-
tical but for their respective means, it follows that Y; also has the same
constant variance:
E(Y; - µ)2 = E(E'f) = a 2 (2.9)
The systematic component, µ, does not explain the variation in Y, but
only its average level. Consequently, the total variation in Y; equals the
variation in the error term.
The assumption in equation (2.7) states that the various error terms
(E¡; i = 1, .. .,n) are statistically independent of one another. Por example,
the fact that the error term of observation i was large should not influ-
ence the size of any prior or successive error terms. We assume, therefore,
that the data have been generated through random sampling. This assump-
tion is not always valid in practice. Por example, if our sample is a time
series, the error terms may well be autocorrelated; that is:
(2.10)
Consequently, the data generating process does not conform to our
assumption that the data were randomly sampled.
As yet, we made no assumption about the shape of the population distri-
bution. However, as we have seen in the previous section, shape matters
when assessing the usefulness of different kinds of averages. So, do we
not make any assumption about the shape of the distribution of the error
term? In fact, in classical statistics, we certainly do. More specifically, we
Modelling an average 53
add the assumption that the error term derives from a normal distribution.
This assumption, together with equations (2.5) and (2.6), can be written
as follows:
(2.11)
which states that the error terms are normally distributed with mean O
and a constant variance.
This is a strong assumption which in practice we should never take for
granted without scrutinising the data first. For example, the normality
assumption appears to be quite reasonable for the data on the demand
for and recruitment of casual labour in Maputo harbour as shown in
Figure 2.1. (Recall that we can judge the variance of E; from the graph of
Y; since, provided our model of the mean is correct, the two variables
have the same variance.) It would be far-fetched, however, to assume that
overtime payments to manual workers in the harbour are also distributed
normally. Figure 2.2 throws serious doubt on such an assumption.
Let us now look at the properties of the sample mean as an estimator
of the population mean, subject to assumptions (2.1) and (2.5)-(2.7).
Thereafter, we shall add the normality assumption which sets the stage
for statistical inference about the population mean.

2.4 THE SAMPLE MEAN AS BEST LINEAR UNBIASED


ESTIMATOR (BLUE)
To judge whether an estimator is good or bad, we need to be clear on
what we mean by the concept of an estimator and how it differs from
that of an estimate. The short answer is that an estimator is a formula we
use to calculate an estimate. An estímate is a numerical value based on
a sample. For example, the generic formula of the sample mean as defined
in equation (2.12) is an estimator:

(2.12)

Once a sample is drawn, we can compute the sample mean by applying


this formula to the data. This computed value of the sample mean is an
estímate of the population mean.
The estímate produced by an estimator depends on the particular
sample we happen to use. It will vary from sample to sample, giving rise
to a distribution of the estimator. The sampling distribution is the prob-
ability distribution of the estimator which, unlike any particular estimate,
is a random variable. Its outcome is unknown until we draw a particular
sample which allows us to calculate the particular estímate. Each sample
54 Econometrics f or developing eo un tries
will yíeld another estímate. The sampling distribution tells us how such
estimates are likely to vary from sample to sample. For this reason, the
sampling distribution is a fundamental concept of classical statistical
theory, but unfortunately it is not always well understood. The difficulty
wíth thís concept arises because we normally have only one sample at our
disposal and hence our main concern is whether the estímate based on
this particular sample is good or bad. In classical statistícs, however, we
do not ask whether the estímate is good or bad but, rather, whether the
estímate was made by a good or bad estimator. To judge whether an esti-
mator is good or bad, we look at its sampling distributíon in general, and
its average and variance in particular.
If the centre of the sampling distribution of an estimator is exactly equal
to the population mean, we say that the estimator is unbiased. That is,
the estimator does not have a systematic tendency to produce estimates
away from the population mean. Obviously, unbiasedness is a desirable
property that we like our estimators to have. The variation around the
centre has implications for the margin of error (precision) in estimation.
The wider the variation around the centre, the larger will be the margin
of error and, hence, the less precise the estimator will be. While an esti-
mator is either unbiased or not, its precision is generally defined in
comparative terms. An unbiased estimator is more precise than a rival
unbiased estimator íf its sampling distribution has a smaller variance. An
estimator is best if it can be shown to have the lowest varíance among a
class of estimators (say, among all linear estimators).
Unbiasedness and precision are criteria we use to assess the quality of
an estimator and to make comparísons among estimators. As we have
seen, they refer to the sampling dístribution of the estimator which is
the distribution of all possíble estimates that can be produced by the
estimator, given the sampling procedure. But does this mean that we have
to enumerate all the possíble samples we can draw from a particular
population? To do this, we would need to know all the elements of the
real population and, subsequently, enumerate all the possible samples we
can draw from this population. This is obviously not feasible in a real-life
situation.
Instead, usually we derive the moments of the sampling distribution
mathematically from the assumptions of the statistical model (i.e. the
assumptions about the model population and about the sampling proce-
dure ). To do this, we make use of the standard properties of the
mathematical expectations and variances of random variables shown in
Appendix 2.1. The bias of an estimator is measured by the difference
between the mean of the theoretical sampling distribution and the popu-
lation mean we seek to estímate. Hence, in the case of the sample mean
the bias will be given by:
BIAS = E(X) - µ (2.13)
Modelling an average 55
If the bias equals zero, that is the mathematical expectation of the esti-
mator equals the population mean, we say that the estimator is unbiased.
Similarly, precision is measured by the variance or, equivalently, the stan-
dard deviation of the sampling distribution. This standard deviation is
called the standard error of the estimator. It gives us an idea about the
possible margin of error involved in estimating the population mean.

Unbiasedness of the sample mean


Given the assumptions of the statistical model (2.1), it can be shown math-
ematically that the sample mean is an unbiased estimator of the population
mean. In other words:
E(Y) = µ (2.14)
Let us take a real-life example to illustrate this property of the sample
mean more vividly. Take another look at the top panel of Figure 2.1,
which depicts the daily demand for casual labour in Maputo harbour
during the period from March 1980 to June 1981. For illustrative purposes,
let us consider this set of data as our real population. If we now draw,
say, 1,000 random samples, each with 15 observations, from this larger
population, the sample mean will obviously vary from sample to sample.
The histogram of the empirical distribution of these sample means gives
us a good approximation of the sampling distribution of the sample mean

0.3

e 0.2
o
t5
~
u.. 0.1

o
o 100 200 300 400 500 600 700 800 900 1000 1100
Population distribution

0.3

e 0.2
o
u~
u.. 0.1

o
o 100 200 300 400 500 600 700 800 900 1000 1100
Sampling distribution
s1ara•M

Figure 2.4 Comparing the sampling and population distributions of labour


demand
56 Econometrics for developing countries
Table 2.3 Means and standard deviations: sampling versus population
distribution
Population distribution Sampling distribution
Mean 574 574
Standard deviation 155 40

in this case. Figure 2.4 compares this histogram of sample means of 1,000
samples (bottom panel) with the histogram of its population distribution.
As shown in Table 2.3, this ( approximate) sampling distribution has exactly
the same mean as the population distribution, which illustrates the fact
that the sample mean is an unbiased estimator. Note, however, that the
standard deviation of the sampling distribution is much smaller than that
of the population distribution, for reasons we shall now explain.

Minimum variance property of the sample mean


The sample mean is a linear combination of the sample values, hence, it
is called a linear estimator of the population mean. The variance of this
estimator is obtained as follows:

(2.15)
n

= -\ I V(Y); + -..\ I Cov(Y; Y)


n ;~1 n i#j

But the assumption that the error term is not autocorrelated means that
the covariances of each pair of Y; and ~ will be zero. Thus equation (2.15)
can be reduced to the following expression:

-
V(Y) = -
ª2 (2.16)
n

Take another look at Table 2.3. It shows that the standard deviation of
the sampling distribution is significantly smaller than that of the popula-
tion distribution. Equation (2.16) tells us why this is the case. The
difference between the sample and population variance depends on the
sample size. It is easy to verify that the calculated standard deviation of
the sampling distribution, 40, approximately equals the standard devia-
tion of the population distribution, 155, divided by the square root of
the sample size, 15. Other things being equal, the larger the sample, the
smaller the margin of error of our estimates.
Modelling an average 57
It can be shown that the variance of any linear estimator of the popu-
lation mean, say U, will be greater than or equal to the variance of the
sample mean. That is,
2 -
V(U) ~ ~ = V(Y) (2.17)
n

The sample mean, therefore, has the property of least variance among all
linear estimators of the population mean.
Hence, the sample mean is the best linear unbiased estimator (BLUE)
of the population mean. This result depends on the assumptions of the
model. It is useful to reflect carefully on how each assumption was used
to prove that the sample mean is BLUE:
1 We assumed that all sample units Y; come from the same population -
i.e. they have the same mean and variance. The assumption of an equal
population mean is critical to proving that the sample mean is unbiased.
2 In addition, we assumed that all the sample units are independent of
each other - i.e. our sample is an independent random sample. This
assumption is crucial for the sample mean to have the minimum vari-
ance property. If the sample units are not independent of each other,
the sample mean will not necessarily have the minimum variance prop-
erty, though it will still be unbiased. As a result, the precision of the
estimator will be in doubt.
Perhaps we can best end with a final word of warning. We have shown
that to prove that the sample mean is BLUE, no assumption was needed
with respect to the shape of the population distribution. But this should
not lead us to believe that shape does not matter. In section 2.2 we saw
that a mean is not always that meaningful. Indeed, if a distribution is
symmetric and preferably bell-shaped, the mean is at its centre. But if a
distribution is strongly skewed the mean is no more than a balance point
with little further interpretative value. As we shall see in Chapter 3, the
sample mean loses much of its power if the distribution of the variable
in question is strongly skewed or riddled with outliers. The validity of the
normality assumption, therefore, is not a luxury, but quite essential to
ensure the power of the sample mean as an estimator.

Exercise 2.3
This exercise can best be done in the context of a classroom workshop.
The aim is to get a better grip on the concept of a sampling distribution.
To do this, take a particular set of data on a variable such as one of the
variables listed in Exercise 2.1. Consider the data as the real population
for illustrative purposes, and draw a number of random samples of equal
size (say, n = 10, or 15, or 25) from this population. (If doing the exercise
in the classroom each student can draw his or her own sample.) Now:
58 Econometrics for developing countries
1 draw a histogram of the population distribution and calculate its mean
and standard deviation;
2 calculate the mean and standard deviation for each sample drawn from
this population;
3 draw a histogram of all sample means and calculate its mean and stan-
dard deviation;
4 check the relation between the mean and standard deviation of the
population and those of the distribution of sample means;
5 comment on the respective shapes of the distribution.
The latter point is particularly instructive if the population distribution
from which samples are taken is strongly skewed. The histogram of sample
means (an approximation of the shape of the sampling distribution) will
tend to be bell-shaped. This tendency derives from what is called the
central limit theorem in statistics. In short, the sampling distribution of
the sample mean will tend to the normal distribution as the sample size
increases, notwithstanding the shape of the population distribution from
which the data were sampled.
(A note on sampling To do this exercise well, it is advisable to use a
reasonably large number of samples. To prevent the exercise becoming
tedious, especially if doing it on an individual basis, it is best to use a
software package which allows you to (a) generate a random variable; (b)
sort the data base by ordering any variable; (c) calculate summary statis-
tics (means and standard deviations) for any subset of the data; and (d)
draw histograms. If one is available, proceed as follows, assuming Y is the
variable which defines the population distribution: (a) generate a random
variable, R; (b) sort the database with respect to R; ( c) select the first
n ( = sample size) observations of Y which will be a random sample of
the wider population; (d) calculate the mean and standard deviation
of this sample of Yvalues; (e) delete R and start again at (a) to generate
the next sample.)

2.5 NORMALITY AND THE MAXIMUM


LIKELIHOOD PRINCIPLE
Let us now introduce the normality assumption. If Y is a normally distrib-
uted variable, its density function, say f(Y), is specified as follows:
1 (X1l2
f (Y) = 1 e- 2 " )
crh'IT (2.18)
Notation: Y - N (µ,cr 2 )
This distribution has two parameters, µ and cr 2 , which are, respectively, its
mean and its variance. In other words, if we know its mean and its vari-
ance, we have all the information necessary to define a normal distribution.
Modelling an average 59
If a sample is drawn at random from a normal distribution then the
probability that the sample value Y will be between limits a and f3 is given
by the area under the density curve over the interval (a, [3) of the range of
Y, as given by:
~

P(a < Y~ [3) = f f(Y)dY


et
(2.19)

Given µ, a 2 , and a, [3, it is possible to compute the numerical value of


the probability by evaluating this integral numerically. The results of such
computations for a = - oo and various values of f3 are routinely available
for the standard normal distribution with zero mean and unit variance:
µ = O and a2 = l. The table for distribution N(0,1), or 'the z table' is
included amongst the statistical tables to this book (see p. 463). The reader
should verify that 95 per cent of the distribution líes in the range -1.96
to 1.96. The probabilities for other normal distributions can be obtained
from this standard normal distribution table by means of a linear trans-
formation which we shall discuss later when dealing with statistical
inference based on the model.

Maximum likelihood function


For theoretical exercises, we can use the density function to express the
necessary probabilities. If Y¡; i = 1, 2, 3,. . . n, is the ith sample unit in
an independent random sample of size n, then because of independence,
the probability of obtaining such a sample is the product of individual
probabilities for each of the sample units. This is the joint probability of
the sample, i.e. all sample units considered together. If the mean and the
variance of the parent distribution are unknown, the joint probability of
the observed sample Y, where Y= (Y1' Y2' Y3 , . . . Yn), is obtained as a
function L equal to the product of the probabilities of each of the sample
units as follows:
(2.20)
where f(Y j µ, a 2) is given by equation (2.18).
The function L is the likelihood function of the sample. It expresses
the probability of obtaining the sample Y, given that it is drawn from a
normal distribution with mean µ and variance a 2 • This likelihood func-
tion allows us to derive an estimator for the population mean µ. To do
this, we ask what the value of µ is that assigns the highest probability to
obtain the sample we actually have in hand. This principle of estimation
is called the maximum likelihood principie. Statistical theory shows that
maximum likelihood (ML) estimators are mínimum variance estimators
among all estimators. The sample mean is the maximum likelihood esti-
mator of the population mean and is, therefore, the least variance
estimator if the normality assumption is reasonably valid.
60 Econometrics far developing countries
The relative efficiency of mean versus median
To illustrate the superiority of the mean when the population distribution
is normal, let us compare its performance with that of the median. It can
be shown that, if the population distribution is normal, the sample median
is also an unbiased estimator, but it will have a larger standard error than
the sample mean. In other words, the sample median is a less precise
estimator of the population mean of a normally distributed variable. In fact,
for large samples from normal distributions, the sample median has a
variance which is approximately 1.57 ( = 7r/2) times larger than that of the
mean (Mosteller and Tukey, 1977: 17). This implies that the standard error
of the median will be approximately 1.25 times larger than that of the mean.
Let us again use our example of the daily demand for labour in Maputo
harbour to compare the relative performance of the mean and the median
as estimators of the population mean. Figure 2.5 compares the (earlier)
histogram of the sample means of 1,000 samples with 15 observations
each, with the histogram of the medians of these samples. Both distribu-
tions have the same mean, 574, which is equal to the population mean,
but different standard errors: 40 for the mean and 50 for the median.
Although the sample size is small, the standard error of the median is
exactly 25 per cent larger than that of the mean. The mean, therefore, is
clearly the better estimator, as you can readily verify if you compare both
histograms in Figure 2.5.

0.3

e 0.2
o
u~
LL
0.1

o
350 400 450 500 550 600 650 700 750
Mean demand for labour
0.3

e 0.2
·uo
~
LL
0.1

o
350 400 450 500 550 600 650 700 750
Median demand for labour
S laJa™

Figure 2.5 Comparing the sampling distributions of mean and median: demand
for labour
Modelling an average 61
Exercise 2.4
This exercise is an extension of Exercise 2.3. Before you calculated the
mean of each sample, now obtain its median as well. Compare the
histograms of the sample means and sample median and comment on
their relative efficiency. If the population distribution is not approximately
normal, you should find that the mean does not necessarily perform better
than the median.

To sum up, if the normality assumption is reasonably valid in practice,


the sample mean is unbeatable as an estimator of the population mean.
It is not only BLUE, but is also the minimum variance estimator among
all possible estimators. Furthermore, as we have seen in section 2.2, a
symmetric and bell-shaped distribution with thin tails has a clear centre
around which the mass of points concentrates. Mean, median and mode
all coincide and, hence, the centre of the distribution lends itself to easy
interpretation. In this case, therefore, the mean is a powerful tool for the
data analyst. Given these theoretical properties of the mean, it is not
surprising that the mean is so prominent in data analysis. But the power
of the sample mean as an estimator relies on the normality assumption
being reasonably valid in practice for the problem at hand. As we shall
see in Chapter 3, if this is not the case, the mean loses much of its power
and may even yield misleading results.

2.6 INFERENCE FROM A SAMPLE OF A


NORMAL DISTRIBUTION

Standard normal distribution and the distribution of the sample mean


The normality assumption not only gives substance to the choice of the
sample mean as the estimator of the population mean, but it also opens
up the possibility for measuring the degree of confidence we have in
making inferences about the population mean from the sample. In applied
work, the problem we confront is that we generally have only one sample
at our disposal on the basis of which we can calculate a sample mean.
But how close is the sample mean to the population mean? Of course, it
is not possible to give an exact answer to this question. But, given that
the parent distribution is normal, we can state the probability that the
estimator produces outcomes within a specified distance from the popu-
lation mean. To do this, we resort to probability theory.
Probability theory tells us that a linear function of a set of variables,
each of which has a normal distribution, also is distributed normally.
Hence, since the sample mean is a linear function of Y;, i = 1, 2, 3, ... n,
where each Y; has a normal distribution, it follows that the sample mean
as an estimator is also normally distributed. To see this more vividly, take
another look at Figure 2.4. The top panel shows that the distribution of
62 Econometrics far developing countries
the daily demand for labour in Maputo harbour is approximately normal
in shape; the bottom panel shows that the histogram which depicts the
sampling distribution of the sample mean also displays a typical normal
shape, but with a much smaller variance than that of the parent popula-
tion. As we have seen, the sampling distribution of the sample mean when
an independent random sample of size n is drawn from a normal distri-
bution is normally distributed with mean µ and variance a2 /n.
Now, the probabilities corresponding to any arbitrary normal distribu-
tion can be derived from the standard normal distribution by means of a
simple linear transformation. The transformation is given as follows:

y - µ,)
( a/...Jn (2.21)
z =

Since Z is a linear function of Y1' Y2' Y 3 , . . . , Yn, each of which is normally


distributed, it follows that Z is also distributed normally. This transformed
variable Z is called a standardised variable. The advantage is that it has
zero mean and unit variance which can be shown as follows:

E (y -µ,)
a/...Jn
= ...Jn E (Y - µ) = O
O'
(2.22)

V (Y-µ,) n - n
al...Jn = a2 .V(Y - µ) = a2 .V(Y) = 1 (2.23)

consequently, Z has a standard normal distribution.


Now, from the density curve of the standard normal distribution we
know that there is 95 per cent probability that a random observation will
be within the range -1.96 and + 1.96. We can, therefore, write:

Y-µ ) (2.24)
P ( -l.96 ,,;:;; al...Jn ,,;:;; 1.96 = 0.95

Multiplying all terms in this inequality by al...Jn and adding µ yields:

p ( µ-1.96 Tn,,;:;;
(J y,,;:;;
- µ + 1.96 Tn ª) = 0.95 (2.25)

What this means is that, under conditions of repeated sampling, the


probability of the sample mean falling within a distance of (±1.96 a/...Jn)
of the unknown population mean is 95 per cent, given the population
distribution is normal with mean µ and variance a 2 •

Confidence intervals
Our main interest, however, is to make inferences about the population
mean based on the sample mean. Now, if the sample mean is, with 95 per
Modelling an average 63
cent probability, within a certain distance of the unknown population
mean, it fallows that the unknown population mean is within a certain
distance of the sample mean. That is, inequality (2.25) can be rearranged
as fallows:
- a - a
P(Y - 1.96 Tn ~ µ ~ Y+ 1.96 Tn) = 0.95 (2.26)

which gives us a probability statement about the location of the popula-


tion mean: under conditions of repeated sampling, there is 95 per cent
chance that the unknown population mean líes within the boundaries of
±l.96al-Ín of the sample mean. Strictly speaking, this statement pertains
to the sample mean as an estimator, and not to any specific estímate
obtained from a given sample. However, we use these boundaries to
construct what is called a confidence interval. A confidence interval gives
us an interval estímate as distinct from a point estímate (i.e. the calcu-
lated value of the sample mean). The advantage of an interval estímate
is that, unlike a point estímate, it gives us a sense of the margin of error
involved in estimation.
It is not correct, however, to say that the population mean líes with 95
per cent probability within the boundaries of a calculated confidence
interval. Indeed, far any given estímate, the population mean líes either
within or outside the boundaries of its confidence interval. There is noth-
ing probable about this, although in real-life situations we do not know
which is the case. A probability statement about confidence intervals, there-
fare, relates to the estimator, not to the estímate. It states that the estima-
tor we use has, in conditions of repeated sampling, 95 per cent probability
of producing confidence intervals which include the population mean.
To see this point, let us perfarm a simple experiment. Suppose, once
more, that the empirical distribution of the daily demand far labour in
Maputo harbour over the period from March 1980 to June 1981 is the
relevant real population with mean 574 and standard deviation 155. Figure
2.6 gives us, respectively, the sample means and their 95 per cent confi-
dence intervals far the first 20 samples (of 15 observations each) that we
generated to produce the histogram of sample mean in Figure 2.4. The
horizontal line in Figure 2.6 depicts the position of the population mean
( = 574). As you can see, all but one (the tenth sample) of the 95 per
cent confidence intervals include the population mean and, hence, 1 in 20
samples (or 5 per cent) misfired. Obviously, if we were to do this exper-
iment again, the number of samples out of 20 which contain the true
population mean would vary from none, to one, to a few. Under condi-
tions of repeated sampling, 95 per cent of the samples will include the
population mean. Any given sample, however, produces a confidence
interval which either includes the population mean, or does not. It would
be wrong, therefare, to say that there is 95 per cent probability that a
particular confidence interval includes the population mean.
64 Econometrics for developing countries
750 -

~
-
700 -
-
-
- - -
- - -
650 -
~ ~ - -
-
- -
- -
-- -
~

600 - - - - -
- -
574
- - -
550 - - -
500 - - - - - - -
-
-
-
- -
- -
- - - -
450 -
-
-
400 -

350 -

2 3 4 5 6 7 8 9 1o 11 12 13 14 15 1 6 17 18 19 20
20 sample means with 95% confidence intervals
SIa Ta""

Figure 2.6 Confidence intervals of sample means: demand for labour, Maputo
harbour

A striking feature of this graph is that different samples drawn from


the same distribution can give quite different results in term of the esti-
mates they yield. Admittedly, in this case our samples were very small
(only 15 observations), but we should not forget that these samples were
drawn from a population which behaved approximately as a normal distri-
bution. In general, therefore, sampling conditions were favourable for
obtaining good results. While most of the samples produce a sample mean
close to the population mean, we should not overlook the fact that a
significant minority of samples nevertheless give us a different picture.
Obviously, sample 10 is an exception, but samples 7, 13, 14 and 16 also
do not perform all that well. In practice, however, we only have one
sample at our disposal and, unfortunately, we do not have the benefit of
knowing the 'true' population mean.
There is, then, inevitably quite a lot of uncertainty about what infer-
ences we can validly make. This conclusion may be a bit disappointing at
first sight. But, in fact, it is a good thing (instead of a bad thing) to be
warned of the inevitable vagaries of random fluctuations. If you despair
about this, never forget that many people are inclined to jump to general
conclusions based on a single sample without ever having the slightest
notion of how day-to-day chance variations may affect the inferences they
make. It is always better to be forewarned than not to have any clue about
whether there is a problem at all. Statistical reasoning at least warns you
to take randomness (chance variations) seriously. This, in itself, is a posi-
tive (not a negative) thing.
Modelling an average 65
Estimating the population variance
In practice, we cannot use inequality (2.26) to produce a confidence
interval because we generally do not know the population variance, a 2 ,
and therefore we need to estímate it in order to compute a confidence
interval for a given sample. How do we obtain an estimator for the popu-
lation variance? We could again resort to the maximum likelihood
principle to find the formula for a 2 which maximises the likelihood func-
tion. It can be shown that the resulting ML estimator of the population
variance becomes:

-n1 L (Y; - - -
Y)2 where Y
1
= -n L Y; (2.27)

But it can be shown that this sample variance turns out to be a biased
estimator of the population variance. The ML sample variance tends to
underestimate the population variance since the bias involves a factor,
(n-l)!n, which is less than l. Consequently, an unbiased estimator of the
population variance is obtained by multiplying the ML sample variance
by n!(n-l), as follows:
1 -
s2 = n _
1
L (Y;- Y)2 (2.28)

where s2 , the sample variance, is an unbiased estimator of population vari-


ance a 2 • As the sample size increases, the factor n!(n-l) will tend to 1
and the sample variance and the ML variance yield similar results. In
other words, the bias inherent in the ML variance will disappear as the
sample size increases. This is typical for many other ML estimators as
well.

The t-distribution
Substituting s for a in equation (2.21) yields a new variable t, as follows:

t (Y-µ,)
= --;¡:r;; (2.29)

The variable t is no longer a linear function of the sample mean since


the formula for s2 includes the square of the sample mean. Consequently,
the variable t does not have a normal distribution. This implies we cannot
simply replace a with s in inequality (2.26) which defines the confidence
interval, and proceed as if nothing has changed. In inequality (2.26), the
boundaries of the confidence interval were obtained by using the critical
values (i.e. ±1.96) of a standard normal distribution. Hence, that state-
ment was based on the premise that Z is distributed normally. As shown
in Appendix 2.2, the variable t has its own distribution, a t-distribution,
which is a symmetric bell-shaped density curve with somewhat fatter
tails than the normal distribution. The t-distribution model has only one
66 Econometrics for developing countries
parameter, its degrees of freedom (df), equal to n-l in this case. But why
are there n-1 degrees of freedom, and not n, the sample size?
In this case, the degrees of freedom refer to the number of indepen-
dent observations in a sum of squares. Note that the definition of t in
equation (2.29) involves the square root of the sample variance as given
by (2.28). The definition of the sample variance itself features a sum of
n squared deviations from the sample mean. However, although this sum
has n components, only n-1 of these can vary freely because, as we shall
see in Chapter 3, the deviations from the sample mean sum to zero. In
other words, given n-1 components of this sum of squares, the last compo-
nent is no longer free to vary. This explains why there are only n-1, not
n, degrees of freedom.
Statistical tables list the critical values of the t-distribution for different
degrees of freedom. Following the same procedure as we did in the
case of a standard normal variable, the confidence interval for the popu-
lation mean, with unknown population variance, can then be obtained
as follows:

(Y - to.95--f,;, Y+ to.95 -+,;) (2.30)

where t0.95 is the critical value of the relevant t-distribution with n-1
degrees of freedom. Note, however, that, as sample size grows, the t-distri-
bution converges to the standard normal distribution; hence, for larger
samples (say, with n > 100) we can safely resort to the standard normal
distribution, even when the population variance is unknown.

Hypothesis testing
A statistical hypothesis is a statement about the value of a parameter of
the statistical model (in this case, about the population mean). The differ-
ence between a confidence interval and a test of a hypothesis is that in
the former we are trying get an idea about the range of likely values of
the unknown population mean, given a particular sample, while in the
latter we are trying to see how likely the sample is for a given hypothe-
sised value of the population mean. In hypothesis testing, we always
consider two complementary hypotheses which do not overlap but,
between them, exhaust all possible values that the relevant parameter can
take. Why do we need two hypotheses? The reason is that hypothesis
testing involves making a decision on whether or not to reject a partic-
ular hypothesis which we denote by H 0 , the null hypothesis. lf H 0 is
rejected, it means that we effectively accept the alternative hypothesis,
called H 1, which contains all other possible values apart from the partic-
ular value specified in H 0 • Hence, H 1 is always very vague in contrast with
the precise nature of Ha: H 1 simply tells us that any other value is possible
apart from the one hypothesised in Ha.
Modelling an average 67
For example, the summary statistics in Table 2.1 tell us that daily recruit-
ment of casual labour in Maputo harbour fluctuated around 500 in the
early 1980s. Suppose that, in a subsequent period, we draw a sample of
observations on daily recruitment and seek to test whether a recruitment
level of 500 continues to be the average. Our two complementary
hypotheses will then look as follows:
(2.31)
The basic idea then is to test the null hypothesis that the sample is
drawn from a population with mean 500. In this case, the alternative
hypothesis specifies that the population mean can be either greater than
or less than 500. This is what is called a two-tailed test.
But suppose we know that transport activity in Maputo harbour
dropped significantly after the early 1980s and, hence, we do not expect
recruitment levels to average as much as 500 per day. In this case, we
specify the complementary hypotheses as follows:
(2.32)
Note that H 0 and H 1 continue to exhaust all possible values for the popu-
lation mean, since the possibility of µ > 500 has been ruled out. This is
a one-tailed test. If we now replace µ by 500 in the expression for t in
(2.29), it follows that our null hypothesis implies that T has a t-distribu-
tion with population mean 500, i.e.:
-
t = Y -_500
1 - t(n-l) 1·t H 0 is
. true (2.33)
shm
since H 0 specifies that µ = 500. How then do we carry out the test? We have
already noted that hypothesis testing involves making a decision as to
whether or not to accept the null hypothesis. This decision will never be fool-
proof. The reason is that we make such decision under uncertainty due to
the random nature of the error term. Our decisions, therefore, will involve
probability statements, not absolute certainties. Table 2.4 lists the two types
of errors which we may encounter when making this type of decision.
A type I error involves rejecting the null hypothesis when in fact it is
true. The probability of making a type I error, a, is called the level of
significance of a test. We always specify this level of significance clearly
befare we do the test. It is customary to allow for a 5 per cent probability

Table 2.4 Types of errors in hypothesis testing


Decision with respect to H 0
True Correct Type 1 error
H 0 is actually Probability = (1 - a) Leve! of significance = a
False Type 11 error Correct
Probability = 13 Power of the test = (1 -13)
68 Econometrics for developing countries
of making a type I error. A type 11 error involves accepting the null hypoth-
esis when in fact it is not true. The power of a test is given by the proba-
bility of correctly rejecting the null hypothesis and, hence, equals (1-!3),
where !3 is the probability of making a type 11 error. In most cases, it is dif-
ficult to calculate this latter probability precisely since this would require
the alterna ti ve hypothesis to be very precise (equal to a particular value ),
rather than vague. However, it is important to choose the most powerful
test for any given level of significance. For example, if the alternative
hypothesis is such that a one-tailed test is feasible, then this test will have
higher power than one using a two-tailed procedure. Finally, other things
being equal (the inherent randomness of the data; sample size), setting a
higher level of significance will increase the risk of wrongly accepting Ha to
be true when in fact H 1 is true.
Box 2.1 reviews the step-by-step procedure required to carry out a test
of hypothesis. To illustrate the procedure, consider our earlier example
of recruitment of casual labour in Maputo harbour. In 1980-81, recruit-
ment levels ftuctuated in a typical bell-shaped fashion around an average
level of 500. Thereafter, the activity at the harbour fell dramatically. We
would expect, therefore, daily recruitment to decline equally. Suppose

Box 2.1 Step-by-step procedure of hypothesis testing

1 Formulate the null-hypothesis, Ha.


2 Formulate the alternative hypothesis, H 1, in such a way that the
rejection of Ha implies the acceptance of H 1; the specification of
the alternative hypothesis determines whether we are dealing
with a one-tailed or a two-tailed test.
3 Specify the level of significance, a, of the test (usually, 5 per
cent).
4 Specify the appropriate test statistic (which requires that the
underlying assumptions are reasonably valid in practice ).
5 Find the appropriate critica! value of the test statistic from the
relevant statistical tables, taking account of whether the test is
one- or two-tailed.
6 Calculate the value of the test statistic on the basis of the sample.
7 Compare the calculated value of the test statistic with the crit-
ica! value(s) obtained from the statistical tables and accept/reject
the null hypothesis depending on whether or not the calculated
value of the test statistic lies outside the critical boundary. As a
general rule for all statistical tests, if the absolute value of the
calculated value of the test statistic exceeds the absolute value
of the critical value given in the tables then we reject the null
hypothesis.
Modelling an average 69
now that we have a random sample of 30 observations on daily recruit-
ment, drawn in the mid-1980s, which gives us a sample mean of 450 with
a sample standard deviation equal to 90. Is this enough evidence to
conclude that recruitment levels have fallen? Let us proceed step by step
to test this hypothesis.
1 Our null hypothesis is straightforward: Ha: µ = 500.
2 Since we have strong reasons to believe that recruitment levels
remained at most the same and probably declined, a one-tailed test
appears to be appropriate: H 1: µ < 500.
3 We shall use the conventional 5 per cent level of significance.
4 Assuming that the population distribution is approximately normal
with unknown standard deviation, the appropriate test statistic under
the null hypothesis is the t-statistic as given in equation (2.29).
5 The critical value ta. 95 with 29 ( = 30 - 1) degrees of freedom in a one-
tailed test with 5 per cent significance equals -1.699, a negative value
since we are dealing with the left tail of the distribution.
6 Using equation (2.29), we find that t = -3.04.
7 Since the calculated t value líes in the tail (-3.04 < -1.70), we reject
the null hypothesis, Ha: µ = 500, and accept the alternative hypoth-
esis that the population mean is less than 500. Applying the general
rule, the absolute value of the calculated t-statistic (3.04) is greater
than the absolute value of the critical value (1.70).
The t-test outlined above is fairly robust, which means that it remains
reliable even if the assumptions of the model are not fully satisfied. But
this does not mean that anything goes. In general, if the data are reason-
ably symmetrical or near-symmetrical, the t-test will perform adequately.
But, obviously, if the data are strongly skewed or the danger of outliers
is great, the t-test cannot be relied upon to make a valid inference about
the population mean.

Exercise 2.5
The population distribution of the demand for labour in Maputo harbour
is assumed to be (approximately) normal, with population mean 574. Five
random samples of the daily demand for labour were drawn, with 15
observations each. The resulting sample means and standard deviations

Table 2.5 Samples of daily demand for labour


Sample Mean Standard deviation
1 589 173
2 636 168
3 473 195
4 561 123
5 664 163
70 Econometrics far developing countries
are listed in Table 2.5. For each sample test the hypothesis H 0 : µ = 574,
assuming (a) that the population variance is known to be 155; and (b)
that the population variance is unknown. (Use the seven steps outlined
in the box to carry out each test.)

Statistical versus substantive significance


Testing a hypothesis means that we assess its validity in the light of often
considerable uncertainty due to chance variations in the sample. The level
of significance of a test specifies this degree of uncertainty or risk involved
in making decisions of this nature. Remember, however, that the term
significance here is used in a statistical sense, and not in any substantive
sense. As we have seen, the level of significance of a test only refers to
the risk of rejecting a true hypothesis.
There is, however, a tendency in applied work to confuse statistical with
substantive significance or importance. All that is meant by a high statis-
tical significance of a test is that there is very little risk in rejecting the
null hypothesis wrongly. But this does not say anything about the substan-
tive significance of the test. In our example of daily recruitment of casual
labour, given a big enough sample, we might reject the null hypothesis
that the population mean equals 500 on the basis of a sample with sample
mean 490. In statistical jargon we say that the sample mean is significantly
different (given the level of significance of the test) from the postulated
value and, hence, we reject the null hypothesis. But, substantively, does
it matter all that much if the population mean only declined marginally?
Suppose the population mean has indeed declined from 500 to 495. We
may be able to spot the difference statistically, given a big enough sample
which yields greater precision. But the plight of these casual labourers in
terms of access to daily work is unlikely to be very much affected by such
a slight decline. Hence, a difference which turns out to be statistically
significant may not be all that important.
Conversely, suppose we only have a very small sample at our disposal
with mean 420 and a large standard deviation. In this case, lack of statis-
tical precision may not allow us to reject the null hypothesis that the
population mean is 500. The difference between the sample mean and the
postulated population mean, therefore, is not statistically significant in this
case. But we may still be left with the worry that a substantive decline in
average recruitment may have taken place which our data, due to the lack
of precision, fail to confirm. Indeed, a possible drop in average recruit-
ment from 500 to somewhere in the region of 420 will affect the plight
of workers significantly (in the substantive sense ).
This bring us to a second point. We never accept a hypothesis in the
literal sense. A statistical test only leads to an inference as to whether or
not a hypothesis is consistent with the sample data. In the former case
we say that the hypothesis is maintained, while in the latter it is rejected.
Modelling an average 71
Therefore, if the hypothesis µ = 500 is maintained, it does not necessarily
mean that we accept µ = 500 to be correct. All we can say is that this
hypothesis is consistent with the data. Only rejection by a statistical test,
therefore, is a conclusive inference, given the level of significance, about
a hypothesis.
To sum up, hypothesis testing is an important tool in data analysis,
which should, however, be used with caution. This section has shown, first,
that it is important to check whether the assumptions of the model are
(approximately) valid in practice. A test is as good as the assumptions
upan which it is based. If the latter are not satisfied, the test is mean-
ingless, even if it 'looks good'. Second, never confuse substantive signif-
icance with statistical significance. The latter tells us something about the
confidence we have in the inferences we make, but this does not neces-
sarily mean that these inferences are substantively important. Obviously,
a good data analyst always seeks to obtain significant statistical results on
issues which matter substantively. But at times our data are insufficient
to establish conclusive proof which allows us to dismiss a hypothesis. The
fact that we maintain such a hypothesis does not mean it is correct.
Conversely, with a large sample we may be able to detect small differ-
ences and render them statistically significant, although substantively it
may well make little difference one way or the other. Hypothesis analysis,
therefore, should not be used blindly but to help us to draw inferences
on issues which matter substantively.

2.7 SUMMARY OF MAIN POINTS


1 Modelling an average requires us to make assumptions about the shape
of the population distribution. That is, an average, on its own, tells us
very little unless we have a good idea about the shape of a distribution.
2 In empirical work, three kinds of averages are commonly used: the
mean, the median and the mode. If the distribution is unimodal and
symmetric, all three measures coincide. If, furthermore, the distribu-
tion is bell-shaped with thin tails, the mass of the distribution is
concentrated around this typical value. Consequently, if the normality
assumption is reasonably valid in practice, the average (mean, median
or mode) gives us the unambiguous centre of the distribution. The
classical way of modelling a population mean is based on this assump-
tion that the variable is distributed normally.
3 Unimodal skewed distributions are much harder to interpret as far as
their average is concerned. Each kind of average highlights a different
aspect of the distribution: the mode yields its maximum, the mean its
balance point, and the median its middle value. But none of these
measures depicts an unambiguous centre of the distribution due to its
lack of symmetry. In Chapter 3 we shall discuss what to do if our data
derive from a skewed population distribution.
72 Econometrics for developing countries
4 To model an average we assume that the data fluctuate randomly
around a constant population mean. The random fluctuations around
this mean are given by the error terms corresponding to the different
observations in a sample. We assume that each error term has zero
mean and a constant variance, and that the various error terms are
statistically independent of one another (i.e. the data are generated
through random sampling). Finally, we also assume that each error
term has a normal distribution.
5 Given the assumptions of a constant mean with homoscedastic and
statistically independent errors, the sample mean is the best linear
unbiased estimator. Unbiasedness implies that, under conditions of
repeated sampling, the mean of the sampling distribution of the sample
mean equals the population mean. An estimator is best if it has the
smallest standard error (greatest precision) within its class of estima-
tors (in this case, linear estimators). If, furthermore, the normality
assumption prevails, the sample mean is the ML (maximum likelihood)
estimator which is the minimum variance estimator among all estima-
tors. In general, if the population distribution is normal, the sample
mean is unbeatable as an estimator of the population mean.
6 As an example, we have shown that, if the population distribution is
normal, the sample mean will be more efficient (precise) than the
median. In fact, the variance of the sampling distribution of the sample
median will be approximately 1.57 ( = Tr/2) times larger than that of
the sample mean.
7 The normality assumption lays the basis for statistical inference. If the
population variance is known, we can construct confidence intervals
or test hypotheses using the normal distribution. If not, we use the t-
distribution which requires us to estimate the standard deviation of
the population from the sample. Confidence intervals and hypothesis
testing involve statements about the value of a parameter of the model:
in this case, about the population mean. The difference between a
confidence interval and a test of a hypothesis is that, in the former,
we try to get an idea about the range of likely values of the unknown
population mean, while, in the latter, we try to see how likely the
sample is for a given value (hypothesised) of the population mean.
8 Never confuse statistical with substantive significance or importance.
The former refers to the fact that the decision whether to accept or
reject a statistical hypothesis always entails the risk or uncertainty of
rejecting a hypothesis when in fact (but unknown to us) it is correct.
All that is meant by a result being highly statistically significant is that
there is little risk of rejecting the null hypothesis when in fact it is
true. But a result which is statistically significant may not be signifi-
cant in any substantive sense. In hypothesis testing, what matters is to
obtain statistically significant results concerning issues of substantive
importance.
Modelling an average 73
APPENDIX 2.1: PROPERTIES OF MEAN AND VARIANCE
Suppose X, Y are variables, and a, b are constants. Then the following
relationships hold:

E(X) = Expectation (theoretical mean) of variable X

V(X) = Variance of variable X


= E[X - E(X)]2 = E(X2) - [E(X)]2
SD(X) = Standard deviation of X= >J V(X)

E(aX) = aE(X)

V(aX) = a2 V(X)

E(X + Y) = E(X) + E(Y)


Cov(X, Y) = Covariance of variables X and Y
= El[X - E(X)] [Y - E(Y)]) = E(XY) - E(X)E(Y)

Cov(X, Y) = O if X and Y are independent

Therefore E(XY) = E(X)E(Y) if X and Y are independent

V(X ± Y) = V(X) + V(Y) ± 2Cov(X, Y)


= V(X) + V(Y) if X and Y are independent

E(aX + bY) = aE(X) + bE(Y)

V(aX + bY) = a2V(X) + b2V(Y) + 2abCov(X, Y)


Cov(aX,bY) = abCov(X,Y)

APPENDIX 2.2: STANDARD SAMPLING DISTRIBUTIONS


Given the distribution of a random variable Y, the distribution of g(Y),
where g is a function of Y with certain properties, can be found by using
probability calculus. If Y1, Y2, Y 3, ••• Yn constitute a simple random sample
from a population, then each Y; has the same distribution as the distrib-
ution of the population, and they are independent of one another. They
are called independently and identically distributed (iid) random variables.
Based on the assumption of normal distribution of the population, distri-
butions of various functions of the sample (Y1, Y2, Y 3, ••• Yn) have been
worked out. The distribution of any function of the sample, such as sample
74 Econometrics for developing countries
mean, sample variance, is called the sampling distribution. A selection of
these standard sampling distributions is given below:
1 Let Y be a standard normal distribution (zero mean and unit vari-
ance). The distribution of Y2 can be worked out to be what is called
a chi-square distribution with one degree of freedom. The degree of
freedom is the only parameter of the chi-square distribution in this
case:
Y - N(l,O) then Y2 - X2( 1l (A.2.1)
It is a positively skewed distribution, i.e. it has a long tail on the right-
hand side.
2 Let Y1, Y2 , Y 3 , ••• Yn be independent variables and each have a stan-
dard normal distribution (a simple random sample from a normal
population), then:
(A.2.2)
i.e. Y has a chi-square distribution with n degrees of freedom.
3 Let Y1 , Y2, Y 3 , ••• Yn be a random sample from a normal distribution
with mean µ and variance cr2 • The sample mean is:
-
y== -1 ¿Y.
"\'
(A.2.3)
n '
The sample variance is:
s2 == _1_ I (Y; - Y)z (A.2.4)
n-1
and it can be shown that (a) sample mean and sample variance are
independent, and (b) the random variable
(n-1) s2
cr2
(A.2.5)

has a chi-square distribution with (n-1) degrees of freedom.


4 Let Z have a standard normal distribution, S have a chi-square distri-
bution with k degrees of freedom, and Z and S be independent. Then:

T-
-~
z (A.2.6)

has a t-distribution with k degrees of freedom. The t-distribution is


symmetric around zero.
5 It follows from (A.2.2), (A.2.3) and (A.2.4) above that:
-
T-~ (A.2.7)
- sNn
has a t-distribution with (n-1) degrees of freedom.
3 Outliers, skewness and
data transformations

3.1 INTRODUCTION

In Chapter 2, we saw that the sample mean is a powerful estimator of


the population mean if the parent distribution from which the data are
randomly sampled behaves (approximately) as a normal distribution:
symmetric, bell-shaped and thin-tailed. In this case, the sample mean is
not only the best linear unbiased estimator (BLUE) but, as maximum
likelihood estimator, it also has the minimum variance property (among
all estimators). But the sample mean rapidly loses its superiority if the
normality assumption is invalid in practice: particularly, when the parent
distribution is strongly skewed or has heavier than normal tails which
implies that it is prone to produce samples riddled with outliers. This
chapter discusses ways to deal with skewness and outliers in data.
The chapter is structured as follows. Section 3.2 starts from the result that
the sample mean is a least squares summary of the location (centre) of the
data. This is, in fact, a further reason for its popularity. But least squares
estimators are not resistant summaries, meaning that they are sensitive to
the presence of outliers in a sample. Section 3.3 compares mean-based with
order-based summary statistics and argues that the latter, unlike the for-
mer, are resistant summaries. This explains why exploratory data analysis
(EDA) favours the use of order-based statistics when dealing with data for
which the underlying conditions are unknown. In this section we shall intro-
duce sorne basic tools to analyse skewness and outliers in data: the coeffi-
cients of skewness and kurtosis as mean-based summaries on the one hand,
and EDA's five-number summary and box plot, on the other. The section
ends with an example which highlights the importance of paying attention
to outliers when examining the difference between female and male life
expectancies for different groups of developing countries (low, lower-mid-
dle and upper-middle income countries). In fact, the presence of outliers in
this example throws sorne light on the question whether or not there exists
a sex bias against women in sorne low-income developing countries.
Section 3.4 turns to the important issue of how to check whether the
normality assumption is reasonably valid in practice. We review two
76 Econometrics for developing countries
practica! procedures: a quick and dirty exploratory device which involves
comparing mean-based with order-based statistics, and a formal test using
mean-based statistics (the skewness-kurtosis test). These procedures point
in two directions as to what to do when the normality assumption is
invalid. First, the data may be reasonably symmetric and bell-shaped, but
have heavier tails than a normal distribution. Second, the data may reflect
a strong unimodal skew. To <leal with the first problem, this section argues
that it is preferable to use a robust estimator such as the median instead
of the mean to estímate the centre of the parent distribution. A robust
estimator is an estimator which is likely to perform well for a variety of
underlying conditions (i.e. deviations from normality), particularly the
presence of fat tails. Section 3.5 shows how simple non-linear transfor-
mations of the data often enable us to eliminate skewness in the data.
Consequently, data transformations allow us to get more mileage out of
the classical model for estimating a population mean based on the
normality assumption. The basic principie is that, if the assumptions of
the model do not apply to the raw data, they may well be valid for the
transformed data. But then we also need to translate the results of our
analysis with transformed data back to the realm of the raw data. We
show that to do so gives us interesting insights as to the question of
choosing an appropriate average when dealing with skewed (raw) data.
Finally, section 3.6 summarises the main points of this chapter.

3.2. THE LEAST SQUARES PRINCIPLE AND


THE CONCEPT OF RESISTANCE
In Chapter 2, we discussed the statistical properties of the sample mean,
given the assumption of the model. In this section we look instead at its
mathematical properties as a numerical summary of the location (centre)
of a batch of data. A numerical summary always involves fitting a pattern
to the data, leaving residuals in its wake. Hence, in the case of the sample
mean, the residuals, e¡, are the deviations of the sample values from their
mean:
Y;= Y+ e¡ (3.1)
Never confuse the residuals of numerical summaries with the error terms
in our models: the latter are theoretical constructs (random variables)
used to model the data, while the former are the leftovers of the numer-
ical summaries we try out on the data. Given the nature of a fit, the
residuals will abey certain mathematical properties.

Mathematical properties of the sample mean


Two properties of the residuals of a sample mean matter a great <leal in
practice: (a) the zero-sum property, and (b) the least squares property.
Outliers, skewness and data transformations 77
The first property merely states that the sum of the deviations from the
mean equals zero. The proof is straightforward:
L e; = L (Y; - Y) = L Y; - L Y = nY - nY = O (3.2)
which shows that positive and negative residuals cancel each other out.
The fact that the sum of the mean deviations for any variable equals zero
is a useful result to remember for the algebra of least squares regression.

Exercise 3.1
Table 3.1 lists GNP per capita (in $) for a small sample of seven African
countries in 1990. Using these data:
1 calculate the mean and median of the sample;
2 calculate the residuals of the sample mean and check whether they
sum to zero.

The least squares property tells us that the sum of squared deviations
of the sample values from the sample mean will always be less than the
sum of squared deviations from any other arbitrarily chosen value, say c.
Mathematically stated, this means:
L (Y; - Y) 2 ~ L (Y; - c) 2 for any e (3.3)
In other words:
-
cminimum =y (3.4)
Figure 3.1 illustrates this least squares property in the case of our sample
on GNP per capita for seven African countries (Table 3.1). As you can
see, the sum of squared deviations of sample values from different values
for e reaches a minimum when e equals the sample mean.

Exercise 3.2
Calculate the sum of squared residuals for the sample in Table 3.1.

Table 3.1 GNP per capita: seven African economies, 1990


Country GNP per capita
Tanzania 110
Chad 190
Malawi 200
Kenya 370
Guinea 440
Lesotho 530
Zimbabwe 640
Source: World Development Report 1992
78 Econometrics far developing countries

300000

280000

260000

240000

230571

220000
250 275 300 325 350 375 400 425 450
e
Sta Id"'

Figure 3.1 The least squares property

Least squares summaries are commonly used in applied work. In fact,


the principle of least squares has a long track record in statistics. lt was
first formulated in 1805 by the French mathematician Legendre and
refined by the famous German mathematician Gauss (Stigler, 1986: 145-6).
The mean is the simplest application of the least squares principle. In
subsequent chapters we shall see that one of the most common methods
to derive a regression line is also based on the least squares principle.
Like the sample mean, other least squares estimators are also BLUE
subject to a similar set of assumptions of the model.
Given this least squares property, it should then not surprise us that
the most commonly used measure of the spread, the variance, is derived
from this minimal sum of squared deviations from the mean:
s2 = -1- ""' -
(Y - Y)2
L (3.5)
n-1 '
where s, the square root of the variance, is the sample standard deviation.

The concept of resistant summaries


These mean-based statistics, the mean itself and the standard deviation,
have become so popular that we no longer question their usefulness
as numerical summaries. But are least squares estimators such as the
mean and standard deviation appropriate in all circumstances? In fact,
Outliers, skewness and data transformations 79
they are not. The reason is that least squares summaries tend to be very
sensitive to the presence of outliers in the data. Another way of saying
this is that least squares summaries are generally not resistant summaries.
What does this mean? The property of resistance can be defined as follows:
Resistance is a property we would like summary statistics to have. If
changing a small part of the body of data, perhaps drastically, can
change the value of the summary substantially, the summary is not
resistant. Conversely, if a change of a small part of the data, no matter
what part or how substantially, fails to change the summary substan-
tially, the summary is said to be resistant.
(Mosteller and Tukey, 1977: 2)
In other words, a resistant summary is unaffected by erratic extreme values
in the sample. To illustrate this property of resistance, consider once more
our simple example of the GNP per capita of seven African countries.
Let us see what happens to the sample mean if, for example, we include
in the sample Botswana, which has a GNP per capita of $2,040, instead
of Lesotho's $530.
Table 3.2 gives the picture: the mean reacted quite dramatically, from
$354 to $570, but the median remained unchanged. The standard devia-
tion more than tripled: from $196 to $673!

Exercise 3.3
Repeat exercises 3.1 and 3.2 with the data for sample 2 in Table 3.2:
1 check how big the residual of Botswana is in relation to the other
residuals;
2 check the relative size of the squared residual of Botswana in the total
sum of squared residuals.
What do you conclude from this exercise?

Table 3.2 Resistance of mean versus median


Country GNP per capita
Sample 1 Sample 2
Tanzania 110 110
Chad 190 190
Malawi 200 200
Kenya 370 370
Guinea 440 440
Lesotho 530
Zimbabwe 640 640
Botswana 2040
Mean 354 570
Median 370 370
Standard deviation 196 673
80 Econometrics for developing countries
The example shows that the mean is a much less resistant summary
than the median. The reason why the mean is so sensitive to outliers
directly follows from the least squares principie itself (the concept of an
outlier is defined more formally below; for now, think of it as an obser-
vation far removed from the body of the data). Squaring each deviation
to calculate the total sum of squares means that large deviations acquire
a disproportionate weight in the total sum. Since the mean minimises this
sum of squares, an outlier will tend to pull the mean in its direction. Note,
for example, that the mean in sample 2 (Table 3.2) has become more
distant from its median. It has been pulled in the direction of the outlier
(Botswana) to minimise the weight exerted by the outlier's large devia-
tion. In contrast, the median is not subject to this kind of pull since it is
merely the middle value in the ordered sample and hence insensitive to
the distance of values in its tails.
The least-squares property also explains why the mean is sensitive to
skewness in the empirical distribution of the data. The one-sided long tail
of such a distribution will exert undue weight on the total sum of squared
deviations and therefore pull the mean in its direction. Hence, the mean
will exceed the median when the distribution is positively skewed (i.e.
skewed towards higher values), and conversely, the mean will be less than
the median when the distribution is negatively skewed. The tail-end of a
skewed distribution runs over a larger distance and so will exert a stronger
pull on the location of the mean.
In applied work, we like our summaries to be resistant to avoid one or
more outliers dominating the scene and leading us astray. This is particularly
true if we are unsure about the actual shape of the population distribution
from which our data are drawn. A single large outlier can exert great
infiuence on the location of the sample mean and, hence, can be misleading.
For example, sample 2 in Table 3.2 might lead us to conclude that African
countries, on average, nearly qualify as lower-middle income countries
(which, in 1990, had incomes above $600 per capita), just because Botswana
happened to be in the sample. The median, in contrast, shows that this is not
the case and, moreover, reveals more clearly that Botswana is the exception.

3.3 MEAN-BASED VERSUS ORDER-BASED


SAMPLE STATISTICS
When analysing an empirical distribution we are interested in its centre
(location), spread and shape. As to the latter, two characteristics matter
a great deal: symmetry and the presence or absence of heavy tails. To
describe a distribution, therefore, we rely on descriptive measures for each
of these characteristics. Traditionally, this was done using mainly mean-
based statistics, but more recently EDA has stressed the importance of
the more resistant order-based statistics to describe and picture empirical
distributions. Let us look at each in turn.
Outliers, skewness and data transformations 81
Mean-based statistics
You are probably most familiar with mean-based statistics. To describe a
distribution, mean-based statistics rely on four measures: the mean itself
(centre), the standard deviation (spread), the coefficient of skewness
(skewness/symmetry), and the coefficient of kurtosis (heavy or thin tails).
All these measures either are, or are derived from, moments of the distri-
bution. Each of these measures can be defined for a theoretical
distribution or calculated on the basis of a sample (computed for an empir-
ical distribution).
The mean is the first uncentred moment of a distribution. As you know
already, the population and sample means are given respectively by:
µ = E(Y) (3.6)
- 1
Y=-IY (3.7)
n '
The spread of a distribution is measured by its standard deviation, the
square root of the variance which is the second centred (around the mean)
moment of the distribution. The population and sample variances are
respectively given by:
a 2 =E (Y - µ)2 (3.8)
1 -
s2 = - I (Y - Y)2 (3.9)
n-l '
Another measure of the variation of a probability distribution is its
coefficient of variation, CV. This is a relative measure defined as the ratio
of its standard deviation to its mean, as follows:
CV =alµ (3.10)
while the corresponding sample estímate is given by:
-
cv = s /Y (3.11)
We now come to the measures of shape - the coefficients of skewness
and kurtosis - with which you may be less familiar. These are derived
from the higher centred moments of a distribution. Let us start with the
coefficient of skewness. For a theoretical probability distribution, this
coefficient, a 3, is derived from its third centred moment (hence, the
subscript 3), as follows:

(3.12)

where a 3 = O if the distribution is symmetric. A distribution is skewed to


the right (meaning its long tail is to the right) if a 3 > O, and to the left
if a 3 < O. This explains why we say that a distribution is positively or
negatively skewed. To see this, take another look at equation (3.12). Its
82 Econometrics for developing countries
numerator features cubic powers of the deviations of Y values from the
population mean. Cubic powers preserve the sign of an expression but
inflate the larger deviations proportionally much more than smaller devi-
ations. If the distribution is symmetrical, negative and positive cubic
powers will cancel each other out. If the distribution is skewed, however,
the long tail will dominate the overall sign of the expression. The cubic
power of the standard deviation in the denominator is used to standardise
the measure and so remove the dimension (i.e. it will not depend on the
units in which the variable is measured).
With sample data, we compute the sample coefficients of skewness, a3 ,
as follows:

(3.13)

The coefficient of kurtosis, cx 4 , of a theoretical distribution is derived from


its fourth centred moment (hence the subscript 4) and defined as follows:
J.l4 E(Y - µ,)4
a4 - --
- ª4 - (3.14)

an expression which measures the 'heaviness' of the tails of the distribu-


tion. The fourth power of the standard deviation in the denominator
standardises the measure and renders it dimensionless. Why <loes this coef-
ficient give us a measure of the heaviness of the tails of a distribution?
To see this, note that fourth powers make each sign positive but inflate
large deviations even more than cubic powers or squares would do. The
presence of heavy tails, therefore, will tend to inflate the numerator
proportionally more than denominator. The fatter the tails, therefore, the
higher the kurtosis. But what is a high or low value of the kurtosis?
Here the normal distribution with its thin tails serves as a benchmark.
As you know, a normal distribution has two parameters, its mean and
its variance, which jointly define the distribution. All normal distribu-
tions are symmetrical (hence their skewness equals O) and have a kurtosis
equal to 3. A symmetric distribution with cx 4 > 3 has heavier tails than a
normal distribution. For example, the t-distribution we carne across in
Chapter 2 has heavier tails than the normal distribution and hence its
kurtosis exceeds 3 (the exact size depending on the number of degrees of
freedom). In contrast, a rectangular distribution which has a body but no
tails has a kurtosis cx 4 = 1.8 which is markedly smaller than that of a
normal distribution.
The sample coefficient of kurtosis, a4 , is obtained as follows:

1_ ~(Y- Y)4
n "'-' '
s4
(3.15)
Outliers, skewness and data transformations 83
In practice, the main purpose of calculating the sample kurtosis is to check
whether a symmetric distribution behaves approximately as a normal
distribution. There is not much point, therefore, in calculating a kurtosis
of a skewed distribution. A unimodal bell-shaped empirical distribution
with skewness close to O and a sample kurtosis close to 3 can be taken
to behave similar to a normal distribution. For example, in Chapter 2 we
made frequent use of the empirical distribution of the demand for casual
labour on the day-shift in Maputo harbour during the early 1980s. The
skewness and kurtosis of this empirical distribution are, respectively,
a3 = -0.013 and a4 = 3.007. Not surprisingly, therefore, this distribution
behaves very similarly to a normal distribution. An empirical distribution
with a 3 roughly equal to zero and a4 > 3 has heavier tails than a normal
distribution would have, while a4 < 3 indicates thinner tails than normal.

Exercise 3.4
For samples 1 and 2 in Table 3.2 compute, respectively, the sample coef-
ficients of skewness and of kurtosis. How does the presence of an outlier
in sample 2 affect both measures?

Exercise 3.5
In exercise 2.1 you were asked to compute means, medians and modes
for a set of socioeconomic variables. For each of these variables:
1 compute the coefficients of skewness and kurtosis;
2 comment whether the distribution can be taken to be symmetrical or
not;
3 if so, whether is has normal tails or not.
What do your results tell you about whether the mean is a good summary
of the data?

Order-based statistics
The order statistics of a sample of observations Y; (i = 1, 2, ... n) are
obtained by rearranging the observations in order of increasing magni-
tude. We denote the resulting ordered sample as Y(i) (i = 1, 2, ... n) such
that Y(I) < Y(z) < ... < Y(n)' where the bracketed subscripts refer to the
position of each observation in the ordered list. Suppose, for example,
that we have the following sample of five observations: 4, O, -3, 5 and -2.
These are the Y; values, i = 1, ... 5. The ordered sample is obtained as
follows: -3, -2, O, 4, and 5. This ordered list contains the Y(il values. Hence,
while Y 2 = O, Y(z) = -2, since -2 is the second value in the ordered list.
The median, Md, is the middle value of the ordered sample, Y(i) (i =
1, ... n), and, hence, splits the ordered list into two halves; that is, half the
84 Econometrics far developing countries
observations Y(J) líe below the median, and the other half above it. Box 3.1
explains how to obtain the median of a sample. For a theoretical probabil-
ity distribution the median is the centre of probability of the distribution:
50 per cent of the probability mass líe below it, and 50 per cent above it.

Box 3.1 Computing the median and quartiles


Let Y; (i = 1, ... , n) be the ordered list of sample observations
arranged in ascending order. To obtain the median and the quar-
tiles of the sample, proceed as follows:

Median
The location of the median depends on whether the sample size, n,
is e ven or odd. First, compute the median depth = (n + 1)/2:
1 If the resulting value is an integer and, hence, the sample size is
odd, the position of the median is given by (n + 1)/2. For example,
if n = 5 and, hence, (n + 1)/2 = 3, the median is Y 3, the third
value in the ordered list.
2 If the resulting value contains a fraction (0.5) and, hence, the sam-
ple size is even, the median is obtained by calculating the arith-
metic mean of Yn 12 and Y(n/2) + 1. For example, if n = 6, (n + 1)/2 =
3.5, the median is obtained by taking the mean of Y 3 and Y4• Note
that positions 3 and 4 are the nearest integers on either side of 3.5.
Note that the median depth indicates how far we have to count
inwards to encounter the median. This concept of depth is also used
to pinpoint the position of an observation in an ordered list with ref-
erence to the median. Hence, the depth of Y; is obtained by counting
its position upwards from the lowest value if Y; lies below the median,
and downwards from the highest value if it líes above the median.

Upper and lower quartiles


To find the depth of the quartiles on either side of the median, first
obtain the truncated median depth by dropping the fraction 0.5, if
present, from the median depth. Then, compute the quartile depth
= (truncated median depth + 1)/2
1 If the resulting value is an integer, it indicates the depth of the
quartiles on either side of the median. For example, if n = 10,
the median depth will be (n + 1)/2 = 11/2 = 5.5. The truncated
median depth is 5. The depth of the quartiles, therefore, equals
(5 + 1)/2 = 3. To obtain the lower quartile, count upwards starting
Outliers, skewness and data transformations 85

with Y1, the lowest value in the sample. To obtain the upper
quartile, count downwards starting with Ym the highest value in
the sample. In our example, the lower quartile will be the third
value from the bottom of the ordered list; the upper quartile is
the third value from the top.
2 If the resulting value contains a fraction (0.5), the quartiles can
be obtained by averaging the two values whose depths are the
integers adjacent to the computed quartile depth. To find the
lower quartile, start from the bottom of the list; to find the upper
quartile, start from the top. For example, if n = 7, the median
depth will be 4 (which equals the truncated median depth
because it does not contain a fraction). The quartile depth then
equals (truncated median depth + 1)/2 = (4 + 1)/2 = 2.5. The
lower quartile, therefore, is obtained by taking the average of
the second and third values in the ordered list: QL = (Y2 + Y3 )/2.
Similarly, the upper quartile is found by averaging the second
and third values from the top of the list: Qu = (Y5 + Y6 )/2.
Source: Hoaglin (1983)

Quartiles divide an ordered sample into quarters. Box 3.1 explains how
to do this. The median itself is the second quartile. Apart from the median,
we have the lower, Qv and upper, Qu, quartiles. The interquartile range,
IQR = Qu - Qv is the most commonly used measure of spread for arder
statistics. It gives us the range of the middle 50 per cent of the data. The
range of the data is another order-based measure of spread and involves
computing the difference between the extreme values, respectively Xu and
Xv of the sample, where Xu = Y(n) and XL = Y(l)· Unlike the IQR,
however, the range (Xu - XL) is not a resistant summary of spread.

Exercise 3.6
Using Table 3.2, find the median and the upper and lower quartiles for
the following samples:
1 sample 1;
2 sample 1 without Zimbabwe;
3 sample 2;
4 the sample of GNP per capita for all eight countries listed.
(Between them, these four cases cover all possibilities listed in Box 3.1.)

Exercise 3.7
Using the variables selected in exercises 2.1 and 3.5, for each variable find
the median and the upper and lower quartiles.
86 Econometrics for developing countries
To look at the shape of an empirical distribution EDA starts from what
is called the five-number summary of the data: the median, Md, the lower
and upper quartiles, QL and Qu, and the two extreme values, XL and Xu.
The box plot, a graphical display of this five-number summary, shows us
the basic structure of the data: more specifically, it shows the location,
spread, skewness, tail length and outliers of a batch of data (Emerson and
Strenio, 1983: 58). To see how to construct a box plot it is best to look
at an example. Figure 3.2 compares the two box plots corresponding to
the two seven-country samples of GNP per capita of African countries
listed in Table 3.2.
First, take a look at the box plot of sample l. The plot is constructed
with the aid of the five-number summary. A box plot consists of a box
and two tails. The box gives the variation of the middle 50 per cent of
the data: its upper and lower boundaries are, respectively, the upper and
lower quartiles. The horizontal line inside the box indicates the position
of the median. The tails (line segments up and down the box) run to the
most extreme data values which do not qualify as outliers. Box 3.2 gives
a simple procedure to determine which data points can be considered as
outliers. In sample 1 there are no outliers and, hence, the tails run up to
both extreme values of the sample. In contrast, sample 2 features an
outlier, the high GNP per capita of Botswana. To highlight its status as
an outlier, we plot it as an isolated point.

2100
2000
1900 o
1800
1700
1600
1500
1400
1300
1200
1100
1000
900
800
700
600 1
500
400
300
200 1 1
100
o
Sample 1 Sample 2
STaTa"'

Figure 3.2 Box plots of GNP per capita (two samples of seven African
countries)
Outliers, skewness and data transformations 87

Box 3.2 The definition of an outlier


Intuitively, an outlier is a data point which is distant from the main
body (say, the middle 50 per cent) of the data. To measure distance,
EDA uses the IQR, the range of this middle 50 per cent of the data.
A data point Y 0 is considered to be an outlier if:
Y 0 < QL - 1.5 IQR or Y 0 > Qu + 1.5 IQR
respectively, lower and upper outliers.
A data point is a far-outlier if:
Y 0 < QL - 3.0 IQR or Y 0 > Qu + 3.0 IQR
Source: Hoaglin (1983).

Detecting outliers: the sex bias against women


This section has aimed to equip you with tools to analyse skewness and
the presence of outliers in data: the coefficients of skewness and kurtosis,
on the one hand, and the five-number summary and box plot, on the
other. Let us now put these tools to work in a concrete case by investi-
gating the relation between the difference in life expectancy between
women and men, on the one hand, and the wealth of nations (measured
by income per capita), on the other. Since our concern here is with
univariate analysis only, we shall investigate this relation between two
variables by comparing the distributions of female-male differences in life
expectancy for three groups of countries: low, lower-middle and upper-
middle income countries.
On average, women live longer than men. However, the discrepancy in
life expectancy between women and men varies with the wealth of nations
as measured by GNP per capita. More specifically, as wealth increases,
women outlive men longer. Table 3.3 summarises the data.
EDA teaches us that we should never look at summary statistics in
isolation. Indeed, EDA's approach to data analysis is based on the premise
that data should be organised in ways which help you to detect key
features and see the unexpected. Therefore, always use numerical
summaries in conjunction with visual (graphical) displays of the data so
that you are able to look at the data and see the relevance of particular
summaries. In this case, as shown in Figure 3.3, comparative box plots
prove to be powerful devices that enable us to look carefully at the struc-
ture of the data while interpreting summary statistics.
At this point, we suggest that you attempt exercise 3.8 befare read-
ing on. This will allow you to compare your observations on the data
with ours.
88 Econometrics far developing countries
Table 3.3 Difference (female - male) in life expectancy by leve! of wealth
(GNP per capita groupings), 99 developing countries, 1990
Low Lower-middle Upper-middle
Sample size 42 40
Order statistics
XL -3 o 3
QL 2 3 5
Median 3 4 6
Qu 4 5 7
Xu 5 8 8
IQR 2 2 2
Outliers
India/N epa! (-2) o o
Bhutan (-3)
Near-outliers
Pakistan, Bangladesh (-1)
Mean-based statistics
Mean 2.600 4.230 5.650
s 1.940 1.760 1.460
CV 0.750 0.410 0.260
a3 -1.400 0.023 -0.500
a4 4.250 3.20 2.330
Source: World Socioeconomic Data (SOCECON data file)

9
8
7
6
5
4

3
2

o
-1

-2 lndia/NP
Bhutan
-3
-4
Low Lower-middle Upper-middle
sra1a•M
Figure 3.3 Comparative box plots of gender differences in life expectancy
Outliers, skewness and data transformations 89
Exercise 3.8
Take a careful look at both Table 3.3 and Figure 3.3. Write down your
general comments. More specifically, comment on the following questions:
1 Is there a relation between gender differences in life expectancy and
the wealth of nations?
2 How do these three distributions differ with respect to spread?
3 Are the distributions similar in shape, or not?
4 What does the little cluster of outliers in the distribution of low income
countries tell you?
5 What did you learn from this exercise in terms of the relative useful-
ness of mean-based versus order-based statistics?

Clearly, as wealth increases so does the difference between female and


male life expectancy: the median difference doubles from three years for
low income countries to six years for upper-middle income countries, a
result which confirms that women outlive men longer as wealth increases.
The means also indicate this basic pattern, but the respective positions of
the means are more strongly infiuenced by the presence of outliers.
In terms of spread, the IQR is resistant to what happens in the tails
and, hence, picks up the variation of the middle 50 per cent of the data
across samples. An interesting feature of the data is that the IQR remains
remarkably stable as we move from low to upper-middle income countries.
Moreover, in all three cases, the medians are situated exactly in the mid-
dle of the box plots. This result normally indicates that the middle body of
the data is symmetrically distributed. In general, this would be an appro-
priate conclusion to draw, but we need to be a bit careful here. Why? The
reason why we should not jump too quickly to this conclusion is that the
published data on life expectancy are rounded to the nearest integer (for
example, 65 or 66, but never 65.5). Consequently, the difference between
female and male life expectancy is always a (relatively small) integer (1, 2,
3, or -2), never a fraction. Did you notice in Table 3.3 that all arder statis-
tics were small integers? There are no data points between any two adjoin-
ing integers, but only gaps. But with IQR = 2, a data point within the box
will be equal to either the median or the lower quartile or the upper quar-
tile. There is no reason, however, to expect that the lower quartile will have
an equal number of observations as the upper quartile.
This does not mean, however, that the constancy of the IQR from sam-
ple to sample does not convey a real message. Indeed, it is remarkable that
the spread of the data does not increase as the level (i.e. the average dif-
ference between female and male life expectancies) increases from sample
to sample. This is interesting because level and spread often do move
together across comparative samples. The decline in the coefficients ofvari-
ation from sample to sample confirms that here the increase in level does
not go hand in hand with the increase in spread. Note, however, that the
90 Econometrics far developing countries
standard deviations (and, hence, the coefficients of variation) are far more
sensitive to the tails of the distributions. This fact is most obvious in the
case of low income countries. lt should not surprise us, therefore, that its
standard deviation is the highest among the three samples.
But undoubtedly, in substantive terms, the most striking feature of the
data is the unusual cluster of negative outliers in the distribution for low
income countries. Three countries, India, Nepal and Bhutan, along with
two borderline cases (in so far as the definition of outliers is concerned),
Pakistan and Bangladesh, all have in common that the life expectancy of
men exceeds that of women, contrary to the dominant pattern across the
world. This finding has led to an extensive debate in the development
literature as to whether there is a sex bias against women in these coun-
tries (see, for example, Wheeler, 1984; Sen and Sengupta, 1983; Sen, 1985,
and Harriss, 1990).
The lesson we want to draw from this example is that we can get much
more mileage out of the data if we use mean-based statistics jointly with
arder statistics and their graphical display. Por example, it is much easier
to see why the sample of low income countries has a much higher kurtosis
than the other samples once we have taken note of the cluster of outliers
in its lower tail. The kurtosis does not by itself explain this feature in the
data. Similarly, a quick look at the standard deviations of the three
samples may lead us to overlook the remarkable constancy of the spread
of the middle 50 per cent of the data, the IQR, as we move from sample
to sample. Level increases from sample to sample, but the mid spread
does not. More importantly, however, merely to calculate the means and
standard deviations of the three samples may cause us to overlook the
outliers in the sample of low income countries. The point here, however,
is to explain the outliers, not to hide them. This is where comparative box
plots show their power as analytical tools of exploratory data analysis.
They draw our attention both to regular patterns within the data as well
as to the unusual or exceptional. In this case, a set of three to five coun-
tries clearly does not fit the overall pattern. This calls for an explanation
as to why the life expectancy of women is less than that of men in these
countries. Or, more vividly, as Sen put it, where are the missing millions
of women in these countries?

3.4 DETECTING NON-NORMALITY IN DATA


If the normality assumption is reasonably valid in practice, the sample
mean is unbeatable as an estimator of the population mean. But how do
we detect non-normality in sample data? This section discusses two proce-
dures to do this: a quick exploratory check and a formal test. Befare doing
either of these, always take a good look at the data first. You can do this
with a histogram or a box plot. In many cases it will be plainly obvious
that the normality assumption is invalid. Take, for example, another look
Outliers, skewness and data transformations 91
at the histogram in Figure 2.2. It does not require a formal test to see
that this sample does not derive from a normal distribution. Similarly, in
Figure 3.3, the box plot of the difference between female and male life
expectancy for low income countries reveals a heavy lower tail which
clearly violates the normality assumption, although the middle 50 per cent
of the data are quite symmetric. As a rule, use numerical summaries and
graphical displays to get a good grip on your data. This in itself allows
you to weed out cases which clearly violate the normality assumption. If
further scrutiny is required, apply the procedures we shall now discuss.

An exploratory check for normality in data


This two-step procedure is a simple application of the tools developed in
the previous section. It involves comparing mean-based with order-based
statistics to check for normality in data. The procedure involves two steps.
First, we check whether the data are reasonably symmetric and, if so,
subsequently verify whether the tails of the sample distribution are suffi-
ciently thin to warrant the normality assumption.

Step 1: Checking skewness


The first step involves comparing the sample median, a resistant summary,
with the sample mean which is non-resistant. The absence or presence of
skewness is deduced from the location of the mean relative to the median,
as follows:
1 positive skewness mean > median
2 approximate symmetry mean = median
3 negative skewness mean < median
To judge whether the mean is approximately equal to the median, it is
useful to compare the distance between them with the IQR. A graphical
variant of this is to draw the position of the mean in the box plot of the
data. Figure 3.4 gives an example: it features the box plot of the distrib-
ution of GNP per capita for 111 countries in 1990 (from data file
SOCECON). The mean (depicted by the horizontal line) is clase to the
upper quartile of the data, while the median is clase to the lower quar-
tile. If the mean and median differ significantly because of the systematic
skewness in the data, it is best to try out a power transformation of the
data to eliminate skewness. We shall show how to do this in section 3.5.
A word of warning is necessary here. It is possible to come across data
where the mean and median differ sufficiently for us to be inclined to try
out a transformation, yet none is called for because the discrepancy
between mean and median is solely due to the behaviour in one of the
tails of the distribution. In fact, our example on the differences between
female and male life expectancy for low income countries is such a case.
92 Econometrics far developing countries
35000 -
o
30000 -

25000 -
o
o
@
20000 - o
§

§
15000 -
(

10000 -
8000 -
6000 -
4000 -
2000 - 1 1

o-
sra1a•M

Figure 3.4 Mean versus median: GNP per capita

9-
8-
7-

6-
5-
4-

3-
2-
-

o-
-1 -

-2 - lndia/NP
Bhutan
-3 -
-4 -

srata™
Figure 3.5 Symmetric but unusual tail: female-male life expectancy
Outliers, skewness and data transformations 93
Figure 3.5 reproduces the box plot along with the position of the mean.
The reasonably significant discrepancy between the mean and the median
may lead us to decide to transform the data. But if we do this, we over-
look the fact that the middle body of the data is symmetric, as indicated
by the position of the median which cuts the box in halves. As discussed
above, the problem here is not skewness, but the presence of a cluster of
outliers in the lower tail.
To avoid failing to distinguish between genuine skewness in data and
un usual behaviour (outliers) in one of the tails (both of which lead to the
divergence between the mean and the median), always observe whether
the middle 50 per cent of the data also manifest skewness. If they do, a
transformation is called for; if not, our attention should go to the unusual
behaviour in the tail. To make this distinction you can use a box plot to
check the location of the median in relation to the quartiles. Symmetry
requires that the median (approximately) divides the box into halves.
Alternatively, you may use Bowley's coefficient of skewness, bs, which is
a resistant measure of skewness defined as follows:
bs = (Qu + QL - 2 Md) I IQR (3.18)
where bs líes within the range -1 and + l. You can easily verify that if
the median is situated in the middle of the upper and lower quartiles, bs
equals O, indicating symmetry in the middle body of the data. If bs < O,
the middle 50 per cent of the data is skewed to the left (negative skew),
and if bs > O, it is skewed to the right (positive skewness).
To sum up, step 1 checks whether or not the distribution is skewed. To
do this, compare the mean with the median to see whether they diverge
significantly. Compute bs, Bowley's resistant measure of skewness, to
verify whether the middle 50 per cent of the data confirm your conclu-
sion. If the distribution is genuinely skewed, you may decide to try out a
transformation of the data to see whether it is possible to eliminate skew-
ness. If the data are approximately symmetrical (or manifest only slight
skewness), you can proceed with step 2.

Step 2: Checking far heavy tails


To detect whether the sample data display thinner or fatter tails than those
of a normal distribution, we may be inclined to compare the standard
deviation, s, which is non-resistant, with the IQR, a resistant measure of
spread. However, unlike the mean and median, the standard deviation
and the IQR do not measure the same thing even if the underlying distri-
bution is symmetric and undisturbed by outliers. Indeed, the IQR
measures the range of the middle 50 per cent of the (ordered) data, while
the standard deviation is a measure of spread derived from the least
squares property. Consequently, the two measures are not directly compa-
rable. But we can nevertheless make use of the property that, for a normal
94 Ecanametrics far developing cauntries
distribution, a definite relation exists between its interquartile range (IQR)
and its standard deviation a:
IQR
a=-- (3.19)
1.35
Using this theoretical relation of a normal distribution, we can now
define an alternative resistant measure of spread, the pseudo standard
deviation (sp), based on the IQR and comparable with a standard devia-
tion, as follows (Hoaglin, 1985: 426-7):
IQR
s =-- (3.20)
p 1.35
Now, if a symmetrical empirical distribution approximately has normal
tails, its pseudo standard deviation will be very similar to its standard
deviation. If not, the two values will diverge. More specifically, we get:
1 sP < s implies heavier than normal tails;
2 sP =s implies (approximately) normal tails;
3 sP > s implies thinner than normal tails.
Consider two examples. First, take once more the data on the demand
for casual labour in Maputo harbour, as depicted in the histogram of
Figure 2.1. As you know by now, these data behave very much like a
normal distribution. So let us see how they stand up to our simple test.
In this case, s = 155 and IQR = (Qu - QL) = (677 - 476) = 201.
Consequently, sP = 201/1.35 = 149, which is only slightly less than the stan-
dard deviation. For practica! purposes, therefore, this distribution can be
taken to behave as a normal distribution (since, as you already know, its
mean and median are close together).
Second, consider the data on the difference between female and male life
expectancy for low income countries. We have seen that this distribution is
symmetric in the middle, but its mean deviates significantly from its median
due to the un usual behaviour of its lower tail. How <loes it perform on our
simple test? Table 3.3 gives us the relevant information: IQR = 2; hence, sP
= 1.48 < s = 1.94. In other words, the tails of this distribution are much
heavier than those of a normal distribution with a similar IQR.
Consequently, we reject the normality assumption in this case.
In conclusion, if a sample of data passes both tests (steps 1 and 2), we
can safely conclude that the data are drawn from a parent distribution
which is approximately normally distributed.

The skewness-kurtasis (Jarque-Bera) test far narmality


It is also possible to test more formally, using the mean-based coefficients
of skewness and kurtosis, whether data are approximately normal in shape.
It can be shown that:
Outliers, skewness and data transformations 95
Z _ a3 ·Vn d Z _ (a 4 - 3)·Vn ( . )
3 - V6 an 4 - \724 3 21

where both have a standard normal distribution in large samples. The


skewness-kurtosis (Jarque-Bera) test for normality tests the joint hypoth-
esis that a 3 = O and a 4 = 3 (i.e. the values of a 3 and a 4 for a normal
distribution). The relevant test-statistic is (Z32 + Z 42 ) which follows a chi-
square distribution with two degrees of freedom. Hence,
H 0: a 3 = O and a4 =3
H 1: a 3 ~ O or a 4 ~ 3 or both (3.22)
Z 32 + Zi = a 32 • (n/6) + (a 4 - 3)2 · (n/24) - X(z)

where the critica! value of the chi-square statistic with two degrees of
freedom equals 5.99 at the 5 per cent level of significance. Strictly
speaking, this test is only valid for (very) large sample sizes: n > 1,000.
The reason is that Z 3 and Z 4 cannot be taken to be independent if the
sample size is smaller than 1,000 and, therefore, the use of the chi-square
distribution as the sampling distribution of (Z32 + Zi) is not strictly valid.
However, the test is still worthwhile asan indicative exercise if our sample
size happens to be smaller, which will often be the case in applied work.
More specifically, if the test rejects the normality assumption, you can
be reasonably assured that the data are not derived from a normal distri-
bution. But if the test does not reject the null hypothesis, you still need
to be careful about its interpretation. It is possible that the test has missed
out one or another indication of non-normality. This apparent ambiguity
should not trouble you too much. The mechanical application of tech-
niques or tests by themselves is never fruitful. Applied data analysis
invariably involves informed judgement along with technical expertise. It
is always useful to look at data in different ways to assess the reason-
ableness of the assumptions involved in modelling. In this case, use the
skewness-kurtosis test in conjunction with the more informal exploratory
checks obtained by comparing resistant and non-resistant measures of
level and spread to judge whether the normality assumption is likely to
be valid in practice.

Exercise 3.9
Using the variables selected in exercise 2.1, 3.5 and 3.7, for each variable:
1 summarise mean-based and order-based statistics;
2 find the outliers, if any, in the sample;
3 produce the relevant box plot;
4 test the normality assumption by (a) using the two-step exploratory
check and (b) using the skewness-kurtosis test.
96 Econometrics for developing countries
Heavy tails: mean versus median revisited
So what do we do if our data are skewed or have heavy tails? If the data
are skewed, a transformation of the data may help. We discuss how to do
this in section 3.5. But what to do if the data are reasonably symmetric
but have heavier than normal tails? Should we continue to use the sample
mean as the preferred estimator of the population mean?
In section 2.5 we showed that the mean is the superior estimator if the
normality assumption prevails. In those circumstances the mean clearly
outperforms the median: the standard deviation of the sampling distrib-
ution of the median will be about 25 per cent larger than that of the mean.
But what if the underlying conditions of the parent distribution are
unknown and, in particular, if it is likely to have heavy tails? In such
cases, an estimator which performs well under very restrictive circum-
stances but does not do so well under a variety of different conditions is
not much use. It is then preferable to use a robust estimator such as the
median. Indeed:
In non normal distributions with long tails, the relative efficiency of
the median to the mean rises and may become larger than l. We can
now define more specifically what is meant by a robust estímate -
namely, one whose efficiency relative to competitors is high (i.e. seldom
much less than 1) overa wide range of parent populations. The median
is more robust as well as more resistant to erratic extreme observa-
tions, although it is inferior to the mean with symmetrical distributions
not far from normal.
(Snedecor and Cochran, 1989: 136)
Hence, in cases where the empirical distribution is reasonably symmet-
rical but sP is significantly larger than s ( or a4 is well above 3), it is
preferable by far to use the median as the estimator of the population
mean (which, for a symmetrical distribution, is also equal to the popula-
tion median).
If we rely on the sample median as our estimator, it is possible to
construct a conservative confidence interval for the population median
that is valid for any continuous distribution, as follows (ibid.: 136-7). Two
values in the ordered list of sample values serve as the lower and upper
confidence limits. To obtain a 95 per cent confidence interval, we obtain
the positions for the confidence interval by first calculating the values of,
respectively, [(n + 1)/2 ---in] and [(n + 1)/2 +--in], subsequently rounding
down the lower value and rounding up the upper value to the nearest
integers.
To illustrate this procedure, suppose we have an ordered sample of the
following 15 observations:
Sample: 1, ~ 3, 6, 8, 1~ 12, 14, 16, 1~ 19, 2~ 2~ 32, 34
Median: 14
Outliers, skewness and data transformations 97
To obtain 95 per cent confidence interval, we calculate the position of the
approximate limits as follows:
(n + 1)/2 - "1n = (15 + 1)/2 - "115 = 4.3, which rounds down to 4
(n + 1)/2 + "1n = (15 + 1)/2 + "115 = 11.87 which rounds up to 12
giving us the positions of the lower and upper confidence limits. Hence,
the lower limit of the 95 per cent confidence interval is 6, the fourth value
in the ordered list, and the upper limit is 24, the twelfth value in the list.
Note that the distance of the lower limit of this confidence interval to the
median is not necessarily equal to the distance from the median to the
upper limit. This will only be approximately the case if the data are fairly
symmetrically distributed. Indeed, since the confidence interval is based
on arder statistics, its limits depend on the shape of the distribution.

3.5 DATA TRANSFORMATIONS TO ELIMINATE SKEWNESS


Many socioeconomic data, when unimodal, are strongly skewed: usually,
but not always, to the right. One reason is that most socioeconomic data
have a clear floor but no definite ceiling. For example, production cannot
be negative, but enterprises differ widely in the level of their output.
Another reason is that socioeconomic data reflect inequalities in the
economy and society. Income distribution, for example, is usually heavily
skewed to the right: most people have relatively low and modest incomes
while fewer and fewer persons have ever larger incomes. Similarly, a time
series which grows at roughly a constant rate of growth will produce a
set of data points which are skewed to the right because the increment
between successive points increases as the level goes up, due to the
constancy of the growth rate.
Skewness in data is a majar problem when modelling an average. First,
as we have seen, a skewed distribution has no clear centre. The mean, its
centre of gravity, will differ from the median, its centre of probability.
Second, the sample mean is no longer an attractive estimator of the popu-
lation mean since the normality assumption is not satisfied. It appears,
therefore, that the model discussed in Chapter 2 is of no use here. Or is
it? Fortunately, it is often still possible to use the powerful properties of
the classical model based on the normality assumption even when working
with data which derive from a skewed distribution. We do this by changing
the shape of the underlying distribution using a non-linear transformation
of the data. This section shows how to do this.
A transformation is a conversion of the data by mathematical means.
The conversion of degrees Celsius into Fahrenheit is an example of a
linear transformation. A linear transformation only changes the scale and
the location of the data, not its shape. To change the shape of the distri-
bution of the data, we need non-linear transformations, for example, a
98 Econometrics far developing countries
power, a square root or a logarithm. The latter, the logarithmic transfor-
mation, is highly popular in applied data analysis. So let us start with this
transformation.

The logarithmic transformation


Logarithms have the property of shrinking the distance between two or
more values which are greater than l. Take a look at Table 3.4 which lists
the GNP per capita of three African countries. The raw data show that the
GNP per capita of Zimbabwe is relatively clase to that of Tanzania, while
Botswana's is clearly way out. Using logarithms, however, shows a differ-
ent picture. The position of Zimbabwe is now closer to that of Botswana
than to that of Tanzania. Why is this? The reason is that logarithms shrink
larger numbers more dramatically than they do smaller numbers, as is
evident from Table 3.4. Consequently, in terms of logarithms, Botswana
is no longer an outlier. This simple example suggests that we can use the
logarithmic transformation to shrink the right tail of a positively skewed
distribution proportionally much more than any of the smaller values in
its main body or its left tail. In this way, a skewed distribution may be
rendered more symmetrical.
Let us try out this idea with a larger sample of data. Figure 3.6 shows the
histogram of the distribution of per capita household income for a sample
of 197 households in Hebei province of China, obtained from the World
Fertility Survey, 1985. The data are available in the CHINA file. As you
can see, this distribution is strongly skewed to the right. What happens if
we make a histogram of the logarithms of per capita household incomes?
Figure 3.7 shows the picture. As you can see, the skewness of the distribu-
tion of household income is drastically reduced after the transformation.

Exercise 3.10
Table 3.5 gives you selected summary statistics for both distributions
depicted in Figures 3.6 and 3.7: respectively, per capita household income
and the logarithm thereof. Check whether either or both distribution
approximately satisfies the normality assumption, using: (a) the two-step
procedure involving the comparison of mean-based and order-based statis-
tics; and (b) the skewness-kurtosis test for normality.

Table 3.4 The shrinking effect of logarithms


Country Per capita Difference from LogiGNP) Difference from
GNP (US$) previous country previous country
Tanzania 110 4.70
Zimbabwe 640 530 6.46 1.76
Botswana 2040 1400 7.62 1.16
Outliers, skewness and data transformations 99

0.56

e
.Q

~
u..

0.00
10.00 1200.00
Per capita household income
stata'M

Figure 3.6 Household income data

0.238579

e
.Q

~
u..

o
2.30259 7.09008
Lag per capita household income
s1a1a"'
Figure 3. 7 Log household income
100 Econometrics for developing countries
Table 3.5 Summary statistics of per capita household income
Summary statistics Per capita Logarithm of per capita
household income household income
Median 153.33 5.03
Mean 198.25 4.95
IQR 166.67 1.10
Standard deviation 170.65 0.89
Skewness (a 3) 2.25 ---0.48
Kurtosis (a 4 ) 10.88 3.17
Sample size 197 197

You will have noted that the logarithmic transformation somewhat over-
corrected the skewness in the original data. Indeed, the transformed data
show a slight but significant skew to the left. Strong positive skewness in
the raw data, therefore, has been tumed into mild negative skewness with
the transformed data. This is the main reason why the skewness-kurtosis
test, at 5 per cent significance level, rejects the null hypothesis that the
data are derived from a normal distribution.
As far as the problem of fat tails is concemed, our log transformation
did a splendid job. You will have found that the resulting pseudo stan-
dard deviation is slightly less than the standard deviation, while the
coefficient of kurtosis is slightly above 3. For all practica! purposes, the
log transformed data have thin tails. Given these results - slight but signif-
icant negative skew and thin tails - you might rightly be inclined to decide
that the log transformation did a satisfactory job. The parent distribution
may not be wholly normal, but this does not mean that the classical model
based on the sample mean as estimator is likely to perform badly. In other
words, the underlying distribution of the log transformed data is unlikely
to be so far away from the normal distribution as to affect seriously the
relative efficiency of the sample mean. Furthermore, for reasons which
will become more apparent throughout this book, the logarithmic trans-
formation is very attractive and, if it is likely to perform reasonably well,
you may feel inclined to stick with it. But, altematively, you may want
to search for a better transformation which corrects for positive skewness
in the data without overdoing it. Which other types of transformations
can we use?

The ladder of power transformation


In fact, EDA uses a graduated family of transformations, the power
transformations, to convert a range of unimodal distributions. These trans-
formations can change both the direction (negative or positive) and the
extent of skewness in the data. Table 3.6 illustrates the hierarchy of these
power transformations and their impact on the skewness in the data.
Outliers, skewness and data transformations 101
Table 3.6 Ladder of powers to reduce skewness
Power p Transformation Effect on skewness
3 Reduces extreme negative skewness
2 Reduces negative skewness
1 Leaves data unchanged
o Reduces positive skewness
-1 Reduces extreme positive skewness

In general, the power transformation is of the form yP, where pis a non-
zero real number. The choice of the power p depends on the nature of
skewness in the original distribution. The higher the value ofp above unity,
the greater is the impact of transformation in reducing negative skewness.
Similarly, the lower the value of p below unity, the greater its impact in
reducing positive skewness, except when p is exactly zero. If p = O, yP = 1
for any non-zero value of Y. Obviously, this type of transformation does
not serve any purpose since all the information about Y would be lost. But
it so happens, however, that the logarithmic transformation fits nicely into
this position of the ladder of transformations and so, by convention, p = O
is the log transformation. The reason is that it reduces positive skewness
less than a negative power (p < O) would do, but more than any power p,
such that 1 < p < O. Thus we obtain a hierarchy of powers, p, in terms of
its effects on reducing skewness. This hierarchy is depicted in the so-called
ladder of powers as shown in Table 3.6.
The power used in transformation need not be only an integer but can
contain fractions as well. Hence, for example, it is possible to use a square
root (p = 0.5) which corrects for milder positive skewness in data.
However, the idea is not that yo u should try to find the exact power,
correct to so many places of decimal, like 0.627 or sorne such number, to
get a perfect symmetry. In practice, a suitable rounded power such as 0.25
or 0.75 will suffice as long a reasonable symmetry is achieved.
Let us now go back to our example of per capita incomes of Chinese
households, depicted in Figures 3.6 and 3.7. The effect of the logarithmic
transformation was somewhat stronger than necessary, resulting in a mild
negative skewness in the transformed data. To avoid over-correcting skew-
ness in the original data, we need to move up the ladder a bit: sorne power
in between O and 1 should do the trick. Por example, we could try a square
root, p = 0.5, or a fourth root, p = 0.25. With a bit of trial and error we
settle on the fourth root: p = 0.25. Table 3.7 gives us the summary statis-
tics and Figure 3.8 shows the histogram.

Table 3. 7 Fourth root transformation of per capita household income data


Power Median Mean IQR Standard Skewness a3 Kurtosis a4
transformation deviation
Fourth root 3.52 3.53 0.95 0.75 0.15 2.98
102 Econometrics far developing countries
0.243655

e
o
~
u.

o
1.77828 5.88566
4th root per capita household income
srata""
Figure 3.8 Fourth root transformation

Exercise 3.11
As in exercise 3.10, check whether the underlying distribution of the fourth
root transformation of the data on household income per capita can be
taken to be approximately normal in shape.

You should find that the fourth root transformation brings the data in
line with the normality assumption. In this case, the fourth root is a super-
ior transformation to the logarithmic one. But for reasons which will
become clear in subsequent chapters, economists often find it easier to
work with a variable log(Y) in an equation rather than with Yº· 25 • Hence,
the choice of an appropriate transformation often involves a trade-off
between one which is ideal for the purposes of data analysis and one
which performs reasonably well on this count but also has the advantage
that it lends itself to a more straightforward interpretation (in substantive
terms) of the results.

Modelling with transformed data


So what have we achieved now in terms of modelling? Clearly, the clas-
sical model of inference about the population mean:
Y;=µ+ E¡ (3.23)
Outliers, skewness and data transformations 103
where the error term is assumed to be distributed normally, is inappro-
priate to model the population mean of the distribution of per capita
incomes of Chinese households. But what about modelling the trans-
formed data? Our analysis suggests two models which are likely to perform
well, namely:
Y/~=µ'+ E¡ (3.24)
and
ln(YJ = µ"+E; (3.25)
where, respectively, µ' is the population mean of Y/• (i.e. µ' = E(Y//,))
and µ" is the population mean of ln(Y;) (i.e. µ" = E(ln Y;)). The fourth
root specification is most appropriate, given the normality assumption of
the classical model. The logarithmic specification is likely to perform
reasonably well inasmuch as the deviations from the normality assump-
tion are unlikely to affect the relative efficiency of the sample mean (vis
a vis, far example, the median) too adversely. Hence, in both cases, we
can take advantage of the attractive properties of the sample mean as an
estimator.
At this point you may be inclined to ask what good it does if the fourth
power or logarithmic transformations render our statistical analysis more
amenable. After all, what interpretation are we to give to the fourth power
of household income, orto its logarithm? The question now arises of how
we go back to the original data. In other words, how does this µ' and µ"
relate to the average of the original data of per capita household income?
The answer lies in the nature of transformation itself. To go back to
the original data, we use the inverse transformation. Let us illustrate this
point with the logarithmic transformation. For example, if 4.70 is the
natural logarithm of per capita GNP of Tanzania, the original income
º
value is obtained by simply raising 4.70 to the power of e: that is, e4·7 =
110. Hence, in the case of the logarithmic transformation we get:
Logarithmic transformation of Y to W: W = loge(Y)
Inverse transformation of W to Y: Y = ew
This application of the inverse transformation of the transformed data
therefore takes us back to the original data. The implication is that no
information is lost in the process of transformation but the point is to
interpret it correctly. If we estímate the mean of log transformed data,
what does this say about the location of the centre of the distribution of
the original data?
To answer this question, let us proceed step by step. First, keep in mind
that if the logarithmic transformation proves to be successful in elimi-
nating skewness, the distribution of the log transformed data will be
(near-)symmetric. Hence, given the (near) symmetry of the distribution,
the mean of the transformed data will be clase to its median.
104 Econometrics far developing countries
Second, the logarithmic transformation alters the shape but not the
arder of the data ( the latter property is the definition of a monotonic
transformation). More formally:
If Y¡ > Y¡ then loge (YJ > loge (~) (3.26)
lt follows that the sample unit which is the median of the transformed
data will also be the median in the original data. Obviously, if the sample
size is even and, hence, the median is obtained by taking the arithmetic
mean of the two middle values, the inverse transformation (i.e. the antilog-
arithm) of the median of the transformed data will diverge somewhat from
the calculated median of the original. In conclusion, the antilogarithm
of the median of the transformed data will be equal to (if the sample
size, n, is odd) or near to (if n is even) the median of the original data.
Consequently, since the sample mean and median of the transformed data
will be approximately equal, it follows that the antilogarithm of the sample
mean of the transformed data will be equal to (or very clase to) the
sample median of the raw data.
Third, while the sample mean of the transformed data is an estimator
of the population mean of the transformed data, its antilogarithm is an
estimator of the population median of the original data. Hence, in conclu-
sion, when estimating the population mean of the distribution of the log
transformed data, the inverse transformation of our estímate will give us
an estímate of the median of the distribution of the original data.

Exercise 3.12
Figure 3.4 shows the distribution of GNP per capita (in the file
SOCECON), a distribution which is skewed to the right. Using the data
in the data file:
1 take the logarithms of the GNP per capita data;
2 calculate the mean-based and order-based statistics of GNP per capita
and of log(GNP per capita);
3 check whether log(GNP per capita) is reasonably symmetric and thin
tailed;
4 compute the antilogarithm of the sample mean of the log(GNP per
capita) and compare it with the mean and median of the original data
(GNP per capita).
What do you conclude?

The same argument can be applied to the whole family of power trans-
formations since all these transformations preserve the arder of the data.
The inverse transformation of the sample mean of the transformed
data gives us an estímate of the population median of the original raw
data. For example, the median per capita household income of the data
Outliers, skewness and data transformations 105
on China is 153.3. The mean of the fourth root of the data is 3.526.
The fourth power of 3.526, which is the appropriate inverse transforma-
tion in this case, is 154.5, which is very close to the median of the original
data.
What applies to the point estímate also applies to the interval estímate
(the confidence interval). Hence, after estimatíng the sample mean of the
transformed data and calculating its confidence limits, we can then take
the inverse transformatíon of this sample mean and of its confidence limits
to obtain an interval estímate of the populatíon median of the original
data. It is not valid, however, to calculate the inverse transformation of
the standard error of the sample mean of the transformed data because
the reverse transformatíon changes the nature of the sampling distribu-
tíon of the estimator obtained by reverse transformation.

Once more: mean versus median


The mean is the centre of gravity of a distribution and the median its
centre of probability. If the distribution is symmetrical, the mean and
median will coincide and both measure the centre of symmetry of a distri-
butíon. In Chapter 2 we sung the praise of the sample mean as an
estimator of the population mean if the normality assumption is approx-
imately valid in practice. Under these conditíons, the sample mean is
indeed unbeatable as an estimator of the population mean (which is also
its median as well as its mode). The classical model is a powerful tool
and whenever possible we seek to apply it, provided we can be assured
that deviations from the normality assumption do not invalidate its power.
If the parent distribution of a sample is symmetric but not normal, inas-
much as its tails are too heavy, we have seen that the mean ceases to be
the best estimator since its lacks resistance to outliers as well as robust-
ness to underlying conditions which deviate too much from the normality
assumption. In such cases, it is preferable to estímate the population mean
with the median, a more robust estimator.
If the parent distribution is skewed but unimodal, it is often, but not
always, possible to render it symmetric and, hopefully, normal by applying
a simple transformation taken from the ladder of transformations. If the
transformation works well, we can again rely on the classical model and
use the sample mean of the transformed data as a powerful estimator of
its population mean (which, given the symmetry of the distribution, is also
its median). After estimation, however, we have to go back to the orig-
inal data characterised by a skewed distribution. For the latter distribution,
there is no centre of symmetry and consequently its centre of gravity (the
mean) will diverge from its centre of probability (the median). The inverse
transformation of the sample mean of the transformed data then gives us
an estímate of the population median of the original data, not of its popu-
lation mean.
106 Econometrics far developing countries
3.6 SUMMARY OF MAIN POINTS

1 The sample mean is a least squares estimator. In general, least squares


estimators have desirable properties if the normality assumption is
satisfied. But the least squares property also accounts far the lack of
resistance of the sample mean to outliers and skewness in the sample.
2 To analyse the shape of a distribution (particularly skewness and the
heaviness of its tails) we can use both mean-based and order-based
statistics: respectively, the coefficients of skewness and kurtosis, on the
one hand, and the five-number summary and box plots (along with the
definition of outliers), on the other. The latter, unlike the former, are
resistant summaries which are more appropriate when analysing data
the underlying conditions of which are largely unknown. These
measures are summarised in Table 3.8.
3 To check whether the normality assumption is reasonably valid in prac-
tice, we suggested two procedures. The first procedure is based on a
comparison of mean-based and order-based statistics and involves two
steps. First, check far skewness in the data, by comparing the mean
and the median of the sample. To avoid skewness being due only to
outliers in the tail, Bowley's resistant coefficient of skewness can be
used to check whether the middle 50 per cent of the data are skewed.
Second, to check whether the data display heavier tails than a normal
distribution, compare the pseudo standard deviation, a resistant
measure of spread based on the IQR, with the standard deviation of
the sample. The second procedure involves a formal test - the skew-
ness-kurtosis test - which verifies whether the data are drawn from a
normal distribution (far which the coefficient of skewness equals O and
kurtosis equals 3).
4 If the data are reasonably symmetrical but have heavier than normal
tails, the sample mean may not be the best estimator due to its lack of
resistance and robustness. Robustness of an estimator is the property

Table 3.8 Moment-based characteristics of a distribution


Population Samp/e Measure of Va/ue far
population
Y - N(O,J)
1 -
First moment E(Y) = µ --¡¡IX;=X Centre o
Second moment E(Y - µ)2 = ª2 1 I (X; -X)2 = s2 Spread
n-1
1 -
Third moment _LE (Y - µ) 3 = a 3 ns3 I (X; -X)3 = a3 Skewness (outliers) o
a3
1 -
Fourth moment lE(Y-µ) 4 =a 4 ns4 I (X; -X)4 = a4 Kurtosis (heavy tails) 3
a4

Note:First moment is given as first moment around zero, whereas the remaining are
moments around the mean, or derived from the latter.
Outliers, skewness and data transformations 107
of the estimator to perform well (relative to competitors) over a range
of different underlying conditions. While the sample mean is superior
when the data are drawn from a normal distribution, the sample median
is the more robust estimator and, hence, preferable when the underly-
ing conditions are unknown.
5 If the data are unimodal but skewed, a data transformation is called
for to correct for the skewness in the data. To do this we rely on the
ladder of power transformations which enable us to correct for differ-
ences in the direction of skewness (positive or negative) and in its
strength. Often, but not always, a transformation renders the trans-
formed data symmetric and, hopefully, also more normal in shape. If
so, the classical model of inference about the population mean using
the sample mean as estimator can again be used.
6 After analysis with transformed data it is necessary to translate the
results back to the original data. If the transformation was successful
and, hence, the transformed data are near symmetrical, the inverse
transformation of the sample mean (and of its confidence interval)
yields a point (and interval) estímate of the population median of the
original data, and not of its population mean. Yet the estímate is
obtained by applying the classical model based on the superiority of
the sample mean as an estimator when the normality assumption is
satisfied, to the transformed data.

ADDITIONAL EXERCISES

Exercise 3.13
Demonstrate algebraically that adding observation Xn + 1 to a sample of
n observations will: (a) leave the sample mean unchanged when Xn + 1
equals the sample mean for the first n observations; and (b)
increase/decrease the sample mean when xn + 1 is greater/less than the
sample mean for the first n observations.

Exercise 3.14
Using your results from Exercise 3.9, choose appropriate transformations
for each of your selected variables. Test for normality in each of the trans-
formed data series and comment on your results.
This page intentionally left blank
Part 11
Regression and data analysis
This page intentionally left blank
4 Data analysis and simple regression

4.1 INTRODUCTION

The model of an average discussed in Chapter 2 depicts data as being


composed of a systematic component (the mean) and random variation
around it (the error term). The study of relationships between variables
extends this idea of an average as the systematic component of a statis-
tical model by making the average of the dependent variable conditional
upon the values of the explanatory variables. Hence, in this case we do
not ha ve only one average (i.e. the unconditional average), but a line or
curve of averages of Y, the dependent variable, for different values of the
explanatory variables. This line or curve of averages is called the regres-
sion of Y on the explanatory variables. Put differently, the systematic
component of our statistical model now becomes a function of the explana-
tory variables. Fortunately, as we shall see, many of the principles and
properties encountered when dealing with a simple average can be applied
to regression analysis. Hence, the least squares principle can be logically
extended to obtain BLUE estimators of the population parameters of the
regression line and, similarly, the normality assumption of the error terms
lays the foundations of statistical inference (estimation and hypothesis
testing) in regression analysis as well as extending the application of the
maximum likelihood principle with its desirable properties.
Simple regression means that we only consider one explanatory vari-
able, X. This chapter deals with simple regression as a tool of data analysis:
more precisely, it deals with simple linear regression, which means that
the conditional means of Y for given values of X are situated on a straight
line. As we shall see below, in many cases data transformations make it
possible to extend the reach of simple linear regression to <leal with non-
linear relations between variables. But why consider two variables only?
In fact, most commonly, a variable is deemed to be affected by a score
of other variables. The study of relations between variables, therefore,
typically involves more than two variables. However, our basic principle
is to proceed from the ground upwards. Later, we shall show that multiple
regression which models the relationship of one variable with a set of
112 Econometrics for developing countries
others is essentially arrived at through a hierarchy of severa} simple regres-
sions. An examination of the relationship of one variable with a set of
others, therefore, requires that we have a good understanding of simple
regression. Many of the problems encountered in the practice of multiple
regression analysis can be understood and resolved through the exami-
nation of the component simple regressions.
This chapter is structured as follows. Section 4.2 introduces the concept
of regression as a line or curve of conditional means of Y for given values
of X. Section 4.3 deals with the classical linear regression model and
reviews the properties of the least squares regression line as an estimator
of the population regression. Section 4.5 considers a simple regression in
practice and argues for the need to check carefully whether the assump-
tions of the model are reasonably valid in practice since least squares
regression, like the sample mean, is not resistant. Although it is not our
purpose in this book to <leal with resistant and robust regression, we shall
nevertheless briefty discuss exploratory band regression as a quick median-
based method to obtain a resistant regression curve. We then discuss a
set of simple analytical graphs which help us to detect problems with the
assumptions of the model. Section 4.6 is an interlude: it deals with the
special case of regression through the origin. Section 4.7 discusses the
important concepts of outliers, leverage and inftuential points in linear
regression. Section 4.8 shows how data transformations can help to
linearise non-linear relations between variables, thus extending the reach
of linear regression analysis. Finally, section 4.9 summarises the main
points of this chapter.

4.2 MODELLING SIMPLE REGRESSION


In mathematics, the equation Y = a + b X implies that there is one and
only one value of Y for a given value of X. In data analysis, however, the
relation between two variables is imperfect in the above sense - there
is generally more than one value of Y for a given value of X (Mosteller
and Tukey, 1977: 262). Consider, for example, Engel's famous empirical
law of consumer behaviour. Using a household budget survey of Belgian
working class families carefully collected by the Belgian statistician
Ducpetiaux in 1855, Engel (a German economist and statistician) observed
that the share of household expenditure on food in total expenditure (i.e.
the Y variable) was a declining function of household income (i.e. the X
variable) (Barten, 1985). This is what one would expect: poorer families
spend a higher proportion of their income on food in comparison with
better-off families. But this relationship is not perfect. Families with iden-
tical income do not necessarily spend an equal proportion of their income
on food. Differences in the demographic composition of families and in
consumption habits and tastes will account for differences in food expen-
ditures. In fact, actual budget studies revea} considerable variation within
Data analysis and simple regression 113
Table 4.1 Share of food in total expenditure
lncome group ('000 Shs) Average proportion of food expenditure (%)
Below 1.0 66.6
1.0-1.9 61.5
2.0-3.9 50.4
4.0-5.9 45.2
6.0-7.9 38.0
8.0-9.9 25.6
10.0-24.9 25.5
Above 24.9 18.2
Ali 50.0

each income class with respect to the proportion of household expendi-


ture spent on food. However, on average, the proportion of food expen-
diture in income declines as income increases. Table 4.1 shows this inverse
relation based on data obtained from the household budget survey of
Tanzania in 1969.
This average inverse relation between the share of food in household
expenditures, on the one hand, and income, on the other, is the regression
line or curve. This is an average relation inasmuch as most of the data
points will not be situated on the regression line/curve. Regression analy-
sis, therefore, aims to establish statistical regularities amidst a great many
erratic variations. The regression line, therefore, depicts the systematic
component of a statistical model which also includes an error term that
accounts for the (often considerable) random variations around the line.
In general, we can write the regression of Y on X as follows:
Y= µ(X)+ Ex (4.1)
where, µ(X) is the average of Y for a given value of X such the average
share of food expenditure in household expenditures, in other words, the
conditional average, Ex is the corresponding error term, and E(Ex) = O.
Modelling the regression of Y on X, therefore, implies two steps. First,
we need to consider the question of the specification of µ(X). The regres-
sion of Y on X is nothing but the locus of µ(X) over the range of X. This
locus can take various shapes depending on the nature of the relation
between Y and X. In practice, with the exception of exploratory band
regression (see section 4.5), we try to describe this locus of conditional
means by a suitable mathematical function: say, a straight line or a non-
linear function which can easily be linearised by applying simple data
transformations. This has two advantages: (a) it facilitates statistical
analysis of the model such as the derivation of estimation procedures or
the examination of the properties of the estimators of the parameters of
su ch functional forms (as, for example, the intercept and slope of a line),
and (b) it allows us to give meaningful economic interpretations to the
coefficients (parameters) of these functions. Our main purpose is not so
114 Econometrics far developing countries
much to find the most accurate description of the locus of the regression
points, but rather, to find a function which yields an adequate approxi-
mation of this locus of conditional means.
At times, the choice of the functional forro of the regression may be
suggested by theory. Por example, theory suggests that the average costs
of a firm vary inversely with capacity utilisation. More often, theory does
not give us guidance on the precise functional forro of a relation but may
allow us to narrow down the range of feasible alternatives. But frequently
a researcher will pick a particular functional forro because its coefficients
lend themselves to convenient economic interpretation, such as, for
example, a marginal or average propensity, an elasticity ora growth rate.
Another approach is to let the data decide on the best specification of
the functional forro of µ(X). In this case, a data analyst has to resort to
hints and clues obtained from sample data to arrive at a judgement about
the best shape of the population regression µ(X). These approaches,
theory-guided versus data-instigated choices, are not incompatible, but
often complement each other. We generally start with hints and clues
provided by theory as to the shape of the population regression, perhaps
taking account as well of convenience in terms of the economic inter-
pretation of the coefficients of alternative specifications, and subsequently
we check whether the sample data support our hunches as to the preferred
functional forro. Hence, the choice of the population regression µ(X)
requires us to find a satisfactory description for the average relation of Y
on X on the basis of the sample data. We call this the sample regression.
But modelling a regression curve is not just a question of choosing an
appropriate functional forro for µ(X) in equation (4.1). The second task
in modelling the regression of Y on X is to specify the stochastic nature
of Ex. As we shall see, we do this in much the same way as in the case of
modelling a simple (unconditional) average. The assumptions we make
about the error term lay the basis for statistical inference, i.e. drawing
conclusions from the sample about the population regression.

4.3 LINEAR REGRESSION AND THE


LEAST SQUARES PRINCIPLE
Finding a sample regression consists of fitting a line or curve to the sample
data. This implies that we make an explicit assumption about the shape
of the curve - a straight line, an exponential curve, a quadratic function,
or whatever - that seems reasonable from the point of view of a priori
information, if any, and of the sample data. Subsequently, we fit the curve
or line using one or another statistical method. In regression, the most
commonly used method is derived from the least squares principle. The
simplest shape is of course a straight line which can be readily estimated
using the least squares principle. The two coefficients of a straight line,
the intercept and slope, are easy to interpret. A straight line may not be
Data analysis and simple regression 115
the most appropriate summary of the relation between Yand X. However,
by starting with this simple shape, we often detect its inadequacies from
the data and are, therefore, in a position to move towards a better spec-
ification. As was the case with the sample mean, the least squares
regression line is an attractive (and, in fact, unbeatable) estimator if certain
conditions prevail: that is, if certain model assumptions are reasonably
valid in practice. The model which renders least squares regression
powerful as an estimator of the population regression is the classical
(normal) linear regression model.

The classical linear regression model


In classical linear regression we assume that the population regression
is a straight line within the context of the following set of model assump-
tions:
(4.2)
The move from the general specification in the previous section to the
classical linear model above, therefore, involved two steps. First, µ(X) is
specified as a linear function of X. Hence, once we know the actual values
of 13 1 and 13 2 we can compute the average Y for any given X by using the
linear function (13 1 + 13 2X). Second, V(Ex) is assumed to be the same, irre-
spective of the value of X. Hence, we assume that the error term is
homoscedastic. In applied terms, this assumption of homoscedasticity in
the context of regression means that the variation around average rela-
tionship is assumed to be similar across the range of X: it neither increases
nor decreases with increasing X.
The parameters of the model 13 1, 13 2 and a 2 , therefore, specify the nature
of the average relation between Y and X and the error variation around
it. The analogous sample regression can be written as follows:
(4.3)
where i = 1, ... n (where n = sample size); b1 is the estimator of 13 1; b 2
the estimator of 13 2 ; e are the residuals. The residual, e, is the sample
equivalent of the population error term, E. In fact, e can be seen as the
estimator of Ex, the error term of the population regression model in (4.2).
However, you should not confuse the residual and the error: they are not
the same thing. The general convention adopted in this text is thus to use
Greek letters to denote population values and lower case to denote the
corresponding sample estimate (the mean is an exception to this rule).
Other texts may use a 'hat' C) to denote sample estimates.
To be able to make valid inferences from the sample about the popula-
tion, in addition to the assumptions about population, we also need to spec-
ify the nature of the sampling procedure. We assume that the sample is
obtained through independent random sampling of Y for a given set of fixed
116 Econometrics for developing countries
values of X. Note that X is assumed to be non-stochastic: there is no chance
variation involved in the X values. This is a strong assumption which
implies that X is controlled by the researcher. This assumption is often quite
appropriate for data obtained through experimentation, but not when our
data result from observational programmes. In development research we
generally observe (but do not control) the variation in Y and X, both of
which are subject to chance variations. For the time being, however, let us
remain with this assumption. Later in this chapter, we discuss what hap-
pens if X is a stochastic variable. If X is non-stochastic, it follows that X
and E must be uncorrelated since only the latter is stochastic.
To sum up, if we have a set of paired observations of sample values
(Y¡, X¡), i = 1 ... n, randomly sampled from a wider population, the clas-
sical linear regression model is based on the following assumptions:
1 the population regression is adequately represented by a straight line:
E(Y¡) = µ(X¡) = 13 1 + 132 X¡; ( 4.4)
2 the error terms have zero mean: E( E;) = O, (4.5)
3 a constant variance (homoscedasticity): V(E;) = cr2, (4.6)
4 and zero covariances (no autocorrelation): E( E¡,Ej) = O,
for all i i= j; (4.7)
5 and X is non-stochastic, implying that E(X¡ E;) = O. (4.8)
Nothing has been said as yet about the shape of the distribution of the
error term E¡. In the classical normal linear regression model we shall
make the further assumption that:
6 the error term follows a normal distribution: E; - N(O,cr2) ( 4.9)
As we shall see, these assumptions make the point that least squares
regression has attractive statistical properties as well as providing the foun-
dations for statistical inference based on the least squares regression line.
But befare we discuss these properties and the inferences to which they
give rise in greater detail, let us first have a look at the sample regres-
sion line itself.

Least squares estimation


Given the sample regression (4.3), its least squares estimators b 1 and b2
are obtained by minimising the sum of squared residuals with respect to
b 1 and b2 : where:
L e? = L (Y; - b1 - b2 X;)2
The resulting estimators of b 1 and b2 are then given by:
- -
b 1 =Y - b1 X (4.11)
b = I(X; - X) (Y; - Y)
2 I(X; - X)2 (4.12)
Data analysis and simple regression 117
where
- 1~ - 1~
Y = - 4J Y; and X = - 4J X; (4.13)
n n

are, respectively, the sample means of Y and X.

Mathematical properties of the least squares regression line


The least squares regression line obeys certain mathematical properties
which are useful to know in practice. The following properties can be
established algebraically:
1 The least squares regression line passes through the point of sample
means of Y and X. This can be easily seen from (4.11) which can be
rewritten as follows:
(4.14)
2 The mean of the fitted (predicted) values of Y is equal to the mean
of the Y values:
Let Y; = b1 + b2X; then we have Y= Y ( 4.15)
3 The residuals of the regression line sum to zero:
1~ 1~ ~-"A
- n L. e., = -n L. (Y, - Y.), =Y- Y= O (4.16)

4 The residuals e; are uncorrelated with the X; values, i.e.:


L e;X; =O (4.17)
5 The residuals e; are uncorrelated with the fitted values Y¡. This prop-
erty follows logically from the previous one since each fitted value of
Y; is a linear function of the corresponding X; value.
6 The least squares regression splits the variation in the Y variable into
two components - the explained variation due to the variation in X;
and the residual variation:
TSS = RSS + ESS (4.18)
where
TSS =L (Y¡ - Y)2
ESS = L (Y;-Y)2
RSS = L e;2 = L (Y; - Y;)2
TSS is the total variation observed in the dependent variable Y. It is
called the total sum of squares. ESS, the explained sum of squares,
118 Econometrics for developing countries
is the variation of the predicted values (b 1 + b2 .X). This is the varia-
tion in Y accounted for by the variation in the explanatory variable
X. What is left is the RSS, the residual sum of squares. The reason
why the ESS and RSS neatly add up to the TSS is that the residuals
are uncorrelated with the fitted Y values and, hence, there is no term
with the sum of covariances.
This last property suggests a useful way to measure the goodness of fit
of the estimated sample regression. This is done as follows:
R2 = ESS/TSS (4.19)
where R 2, called R-squared, is the coefficient of determination. It gives us
the proportion of the total sum of squares of the dependent variable
explained by the variation in the explanatory variable. In fact, R 2 equals
the square of the linear correlation coefficient between the observed and
the predicted values of the dependent variable Y, computed as follows:
Cov (Y, Y) ¿(Y; - Y) (Y; - y)
(4.20)
r = Y[V(Y) V(P")] = \/[¿(Y; - Y) 2 ¿(Y;- Y) 2

A correlation coefficient measures the degree of linear association


between two variables. Note, however, that if the underlying relation
between the variables is non-linear, the correlation coefficient may
perform poorly, notwithstanding the fact that a strong non-linear associ-
ation exists between two variables.

Association versus causality


Regression analysis allows us to investigate the statistical assocrnt10n
between two (or more) variables. Association merely means that two vari-
ables covary together, but in itself says nothing about any causal link
between them. The terminology we use in empirical work often suggests
the contrary. Hence, we talk about the dependent and the independent
variables, which suggests that the former depends on the latter. This
usually indicates more than the direction of association: it implies causality
as well. Similarly, we often talk about X as the explanatory variable, which
clearly implies that X is seen as a causal factor of Y. A more neutral but
less vivid terminology consists of referring to the variable X as the
regressor and Y as the regressand. Whatever your preference with regard
to terminology, never forget that a regression model only depicts statis-
tical association between two variables, but in itself it cannot establish the
direction of causality between them. Whether a causal link exists between
two variables and which way the causality runs is a matter which can only
be settled by sound theoretical reftection. This fact should be borne in
mind, although there are various econometric tests of causality (discussed
in Chapter 13).
Data analysis and simple regression 119
Exercise 4.1
The file TPEASANT on the diskette contains data (based on a survey of
600 Tanzanian peasant households) on, respectively, the average size of
agricultura! landholdings, L, and the average household size, H, for eight
groupings by size of land holdings. Use these data to:
1 regress L on H and compute the R 2 ;
2 check whether the mathematical properties listed above hold for this
regression;
3 regress H on L and compute its R 2 •
4 Why does the regression of L on H differ from that of H on L?
5 In your opinion, which regression best depicts the direction of
causality?

The Gauss-Markov theorem: least squares estimators as BLUE


Like the sample mean, the least squares estimators have desirable prop-
erties subject to the assumptions of the classical linear regression model
(i.e. not including the normality assumption). First of all, note that the
least squares estimators b1 and b2 can be shown to be linear functions of
the sample values of Y, just like the sample average.
The Gauss-Markov theorem states that the least squares estimators are
best among the class of all linear unbiased estimators, i.e.:
(4.21)

V(b 1) and V(b 2) have minimum variances among


all unbiased linear estimators of 13 1 and 13 2 (4.22)
In fact, the sample mean is a special case of the Gauss-Markov theorem.
Suffice it to say, however, that the assumptions of homoscedasticity and
lack of correlation of the errors E; are crucial to establish the validity of
the minimum variance property of b1 and b2 •

Maximum likelihood and the normality assumption


lf we now make the further assumption that the distribution of the error
term E has a normal distribution:
E - N(O,a 2 ) (4.23)
we can again show that the least squares estimators derived above also
happen to be the maximum likelihood estimators. Hence, apart from being
BLUE, the least squares estimators turn out to have the minimum vari-
ance property among all unbiased estimators. As was the case for the
sample mean, the least squares estimators of the regression coefficients
are unbeatable if the assumptions of the classical normal linear model are
120 Econometrics far developing countries
reasonably valid in practice. But, as we shall see below, these least squares
estimators are neither resistant nor robust to a wider range of different
underlying conditions. Least squares estimators, therefore, are high-class
performers under ideal conditions when the normality assumption is
reasonably valid in practice. The normality assumption also allows us to
construct confidence intervals and to test hypotheses. This is the issue to
which we now tum.

4.4 INFERENCE FROM CLASSICAL NORMAL


LINEAR REGRESSION MODEL
The treatment of inference in the classical normal linear regression model
is very similar to our discussion of the classical model for estimating a
sample mean. Consequently, in this section we briefly review the main
points without much further elaboration, apart from a few specific points
which concern regression only.

Standard errors
Given the assumptions of the classical linear regression model, the vari-
ances of the least squares estimators are given by:
1 _X2 )
var (b 1) = a2 ( ;;, + ~(X; _ X) 2 (4.24)

a2
var (b 2) =~(X;_ X) 2 (4.25)

Furthermore, an unbiased estimator of a 2 is given by s2 as follows:


~(Y; - b 1 - b2 X;) 2 RSS
s2 = =-- (4.26)
n-2 n-2
where s is called the standard error of regression. Replacing a 2 by s2 in
(4.24) and in (4.25), we get unbiased estimates of the variances of b 1 and
b 2 • Obviously, the estimated standard errors are the square roots of these
variances.
The total sum of squares of X, l (X¡ - X:)2, which features in the denom-
inator of the variances of the intercept and slope coefficients, is a measure
of the total variation in the X values. Thus, other things being equal, the
higher the variation in the X values, the lower will be the variances of
the estimators, which implies higher precision in estimation. In other
words, the range of observed X plays a crucial role in the reliability of
the estimates. Think about this. It would indeed be difficult to measure
the response of Y on X if X hardly varied at all. The greater the range
over which X varies, the easier it is to capture its impact on the variation
in Y.
Data analysis and simple regression 121
Sampling distributions
To construct the confidence intervals and to perform tests of hypotheses
we need the probability distribution of the errors, which implies that we
use the normality assumption of the error terms. Under this assumption,
the least squares estimators b 1 and b2 each follow a normal distribution.
However, since we generally do not know the variance of the error term,
we cannot make use of the normal distribution directly. Instead, we use
the t-distribution defined as follows in the case of b2 :
b2 - Ho(/32)
t= se( b ) - t(n-2) ( 4.27)
2
where H0 (~ 2 ) is the hypothesised value of ~ and se(b 2) is the standard
error of b2 , given by:

(4.28)

using (4.25) and (4.26). The statistic, t(n-2), denotes the Student's t-distri-
bution with (n - 2) degrees of freedom. The reason we now have only
(n - 2) degrees of freedom is that, in simple regression, we use the sample
data to estimate two coefficients: the slope and the intercept of the line.
In the case of the sample mean, in contrast, we only estimated one para-
meter (the mean itself) from the sample.
Similarly, for b 1, we get:
b1 - Ho(/31)
t= se ( bl) - t(n-2) ( 4.29)

where se(b 1), the standard error of b 1, is given by:

[~ + "L(X~ X )]1
12

se(b 1 ) = s 2 (4.30)

Confidence intervals for the parameters f3 1 and f3 2


The confidence limits for ~ 2 and ~ 1 with (1 - a) per cent confidence
co-efficient (say, 95 per cent, in which case a = 0.05) are given by:

b2 ± t ( n-2,~) se(b2) (4.31)

b1 ± t ( n-2,~) se(b 1)
(4.32)

respectively, where t(n - 2, a/2) is the (1 - a/2) percentile of a t-distribu-


tion with (n - 2) degrees of freedom, and se(b 2) and se(b 1) are given by
(4.28) and (4.29) respectively.
122 Econometrics for developing countries
Confidence interval for the conditional mean of Y
At times, we may be interested to construct a confidence interval for the
conditional mean. For example, after fitting a regression of household
savings on income, we may want to construct a confidence interval for
average savings, given the level of income, in arder to assess the savings
potential of a certain type of households. Suppose:
µo = 131 + !32Xo (4.33)
i.e. µ 0 is the conditional mean of Y given X = X 0 . The point estimate of
µ 0 is given by:

b1 + b 2X 0
while its (1 - a) per cent confidence interval can be obtained as follows:

µ, 0 ± t ( n-2, ~) se(µ, 0) ( 4.34)

where its standard error is given by:


1 (X - X)z ]112
se(µ, 0) = s [ -;,, + L:(;; _ X) 2 (4.35)

Confidence interval for the predicted Y values


There are other occasions when we might be interested in the uncertainty
of prediction on the basis of the estimated regression. For example, when
estimating a regression of paddy yield (physical output per unit area) on
annual rainfall, we may want to predict next year's yield given the anti-
cipated rainfall. In this case, our interest is not to obtain a confidence
interval of the conditional mean of the yield, i.e. the mean yield at a given
level of rainfall. Rather, we want to find a confidence interval for the yield
(Y0) itself, given the rainfall (X0). In this case:

(4.36)

where µ 0 is given by (4.31). The (1 - a) per cent confidence interval for


Y 0 given X = X 0 is then obtained as follows:

Y0 = t( n-2,~) se(Y 0) (4.37)

where
- [ 1 (Xo - X)2 ]112
se(Y0 ) - s 1 + -;,, + L:(X; _ X) 2 (4.38)

In this case, therefore, the standard error of Y 0 is larger than that of µ 0 ,


since the latter corresponds to the conditional mean of the yield for a
Data analysis and simple regression 123
given level of rainfall, while the former corresponds to the predicted value
of the yield. In both cases, (4.33) and (4.36), the confidence intervals will
be larger the further away the X value is from its mean in the sample.

Standard error of a residual


Finally, the residuals e¡ are the estimators of errors E¡. The standard error
of e; is obtained as follows:
h = _!_ (X; - X)2
se(e;) = sV(l - h;) where (4.39)
; n + L(X; - X) 2
where s is given by (4.26). Note that while the standard deviation of the
error term is assumed to be homoscedastic, equation (4.37) shows that
the residuals of the regression line are heteroscedastic in nature. The stan-
dard error of each residual depends on the value of h;.
The statistic h; is called the hat statistic: h; will be larger, the greater
the distance of X; from its mean. A value of X that is far away from its
mean (such as an outlier in the univariate analysis of X) will produce a
large hat statistic which, as we shall see in section 4.7, can exert undue
influence on the location of a regression line. A data point with a large
hat statistic is said to exert leverage on the least squares regression line,
the importance of which will also be shown in section 4.7.

Hypothesis testing
The sampling distributions given in (4.27) and in (4.29) can be used for
tests of hypotheses regarding the intercept and the slope of the popula-
tion regression in much the same way as we did in the case of hypothesis
testing concerning the population mean of a univariate normal distribu-
tion. Remember, however, that in this case the t-distribution has (n - 2)
degrees of freedom instead of (n - 1). We shall illustrate the use of the
t-test in the exercises with section 4.5.

What if X is stochastic?
The inferences we make from the sample regression about the parame-
ters of the population regression are contingent upan the assumptions of
the classical normal linear regression model. One of these assumptions
states that X is non-stochastic but a given set of values. As we pointed
out above, this is perhaps a plausible assumption when regression analysis
is done in the context of experiments in which the researcher has control
over the values of X. But in development research our data derive from
observational programmes where neither Y nor X is subject to control.
Consequently, in most cases, both Y and X have to be considered as
stochastic variables.
124 Econometrics for developing countries
How does this infiuence the validity of our inferences? If X is stochastic,
the critica! question turns out to be whether X; and E; are statistically
independent from one another, and consequently do not covary. If:
E(X¡,E;) = 0, for i = 1, ... n (4.40)
it can be shown that the least squares estimators retain their property
of unbiasedness and, furthermore, the variances of the estimators, the
confidence intervals and the tests of hypotheses that we derived in this
section remain valid conditional upan the realised values of X. In other
words, provided X is independent of the error term, once a sample is
drawn (and hence the observed values of X are known), our inferences
are valid contingent upan these 'given' X values.
This may appear rather restrictive. However, if we make the additional
assumption that Y and X are jointly distributed as a bivariate normal
distribution, all formulae derived above with regard to the estimators of
the population regression coefficients, their standard errors, the confidence
intervals and the tests of hypotheses are all valid. In this book, we shall
not go in to detail about the bivariate (or, for that matter, multivariate)
normal distribution. Suffice it to say that if Y and X jointly follow a
bivariate normal distribution, both the marginal and conditional distribu-
tions will be normal, and, importantly, the regression of Y on X is linear.
However, the converse is not true. If we find that Y and X each follow
a normal distribution, this does not imply that they are jointly normally
distributed (see, for example, Maddala, 1992: 104-5; Goldberger, 1991:
68-79).
This completes our brief review of statistical inference in the classical
normal linear regression model. You will undoubtedly have noted the simi-
larities with the problem of inference about the population mean of a
univariate normal distribution. The latter is in fact a special case of the
general regression model. As with the sample mean, the least squares
regression line turns out to be a powerful tool of analysis if the assump-
tions of the classical normal linear regression model are approximately
valid in practice or if Y and X are jointly normally distributed. But if the
assumptions of the model are likely to be invalid in practice, the least
squares line, like the sample mean, rapidly loses its superiority as an esti-
mator of the population regression. Hence, befare embarking on statistical
inferences based on the least squares line, it is important to check care-
fully the validity of the assumptions of the model in practice. This is the
issue to which we now turn.

4.5 REGRESSION WITH GRAPHICS: CHECKING


THE MODEL ASSUMPTIONS
The strength of a statistical model, however powerful it may appear on
paper, rests upan its assumptions being reasonably valid in practice. The
Data analysis and simple regression 125
least squares regression model is a powerful research tool but, as with
most sophisticated tools, when applied inappropriately it easily leads to
nonsensical results being brandished about as deep insights or truths based
on hard facts. This danger is particularly acute now that we have easy
access to powerful computers which allow us to run a multitude of regres-
sions at virtually no cost at ali. Unfortunately, many researchers appear
to switch their heads off when turning the computer on. Often the aim is
to arrive through trial and error at a regression which looks good on
paper, never mind its fragile foundations. Good data analysis, however,
requires that we take modelling and the assumptions upon which it rests
seriously.
In simple regression, the kind of questions we want to investigate so as
to check whether the model assumptions are reasonably valid in practice
are as follows:
1 Does the relation between Y and X follow a linear pattern?
2 Are the residuals reasonably homoscedastic?
3 Are the residuals autocorrelated?
4 Are the residuals approximately normally distributed?
5 Do all data points contribute roughly equally to determine the posi-
tion of the line (i.e. its intercept and its slope ), or do sorne points exert
undue influence on the outcome?
6 Are there any outlying data points which clearly do not fit the general
pattern?
To answer these types of question, EDA teaches us that it is preferable
to combine numerical summaries with graphical displays which focus our
attention on both the regularity of data as well as their unusual, often
unexpected, features. As to graphics, in bivariate analysis, a simple but
powerful tool is the scatter plot which displays either (a) the raw data
themselves (i.e. a plot of Y against X), or (b) quantities derived from the
data (such as, for example, a plot of the residuals against the X values or
against the predicted values of Y). As to numerical summaries, EDA
prefers the use of resistant summaries when the underlying conditions of
the population regression are unknown since the least squares regression
estimators, like the sample mean, lack resistance, as we shall see in section
4.7. In this book, however, we shall deal mainly with least squares regres-
sion, and therefore we do not discuss median-based resistant methods of
estimating a regression line (see, for example, Hoaglin et al., 1983). But
we shall nevertheless make use of exploratory band regression (Hamilton,
1992: 146-7): a simple, but powerful median-based graphical method
of obtaining a regression curve without imposing any functional form on
the shape of the regression. We do not use this method as an alternative
to least squares regression, but as a diagnostic tool to verify whether
our least squares simple regression is a reasonable summary of the data.
The method of exploratory band regression is based on the concept of
126 Econometrics far developing countries
regression as a conditional average of Y for given X. Let us use a simple
example to illustrate its usefulness in applied work.

Exploratory band regression


Take another look at Figure 2.1 (p. 47) which depicts the empirical distri-
butions of, respectively, the demand (D) for and recruitment (R) of casual
labour on the day shift in Maputo harbour during the early 1980s. Both
distributions are approximately normal in shape. Obviously, the two vari-
ables are not unrelated. More specifically, we expect recruitment to
depend on demand. In fact, it is tempting to assume that both variables
are distributed jointly as a bivariate normal distribution since each of the
variables is approximately normal. Suppose we jump to this conclusion
and estimate the regression line of R on D, which yields the following
results (standard errors in brackets):
R = 184 + 0.55 D R 2 = 0.61 (4.41)
(12) (0.02)
As a numerical summary, this result looks quite good: the standard errors
are small and the R 2 indicates that 61 per cent of the total variation in
recruitment can be explained by variations in the demand for daily labour.
But take a look at the scatter plot in Figure 4.1 which shows the data
along with the estimated regression line. You will see a band of observa-
tions lying along a straight line. This line is the 45º line, and so represents
those days on which recruitment equalled demand.
The scatter plot reveals two problems: heteroscedasticity and non-
linearity. Moreover, this example clearly shows that we cannot assume
that two related variables, each of which are approximately normally
distributed, will of necessity jointly follow a bivariate normal distribution.
What went wrong with our initial hunch?
A moment's reflection reveals the problem. In Chapter 2, we saw that
mean recruitment is well below mean demand and, moreover, the stan-
dard deviation of recruitment is significantly lower than that of demand.
At that point, we hinted that it appeared as if the port experiences a
shortage of labour on its day shift when demand is high. This seems to
be the problem. When demand is low, the scatter in Figure 4.1 clearly
shows that recruitment equals demand. However, as demand increases,
recruitment progressively but erratically falls short of demand, indicating
supply shortages. This explains why the plot is heteroscedastic as well as
non-linear.
So what about exploratory band regression? Exploratory band regres-
sion is a simple technique which starts by dividing the scatter plot into a
series of equal vertical bands; subsequently, within each band we find the
point of medians of Y and X; and finally, we connect these points
of medians with straight line segments (Hamilton, 1992: 166). The line
Data analysis and simple regression 127
784 o
o
o o

o o
o
o o
o
o
o o
o
o

124 o o

124 1010
Demand for labour: day shift

Figure 4.1 Regressing D on R

segments which connect these points of medians of the successive bands


trace a curve which approximates the regression of Y on X without
imposing any a priori shape on this curve. Figure 4.2 gives the exploratory
band regression of R on D using four bands. The curve connecting the
points of medians of these four bands clearly reveals that the underlying
regression is non-linear: as D increases, the slope of the curve, while
remaining positive, declines significantly.
As a diagnostic tool, exploratory band regression helps us to trace the
nature of the non-linear relation between Y and X. Furthermore, it is
resistant to the pull of data points which, because of their location, may
exert influence on the regression line. The resistance of the exploratory
band regression follows from the fact that it is based on medians, and not
on means. As explained above, exploratory band regression is not really
an alternative to least squares regression since it does not involve fitting
any particular functional form to the data. In fact, the curve it traces by
connecting successive -medians depends on the number of bands we care
to consider. As such, exploratory band regression <loes not give us a set
of coefficients which allow for straightforward economic interpretation. In
this sense, therefore, exploratory band regression is a diagnostic tool and
not an alternative to fitting a line to the data. But it helps us to choose
which line to fit.
128 Econometrics far developing countries
784

.i:::
, . . ·.
:.een ~<"'-. •.
,,1 ·.... ··.
¡u-
"O ...<<'· . . ".....·~ :: ~ ~· : ·... :·
•.

/~('.i¡:/ ,·.
~
o
.e
.E!!
o
e:
Q) ~/. ,•'•

É
·s .· ,·'
u
Q)
a:

124
124 1010
Demand for labour: day shift
s1a1a™
Figure 4.2 Exploratory band regression: D on R

Scatter plots and residuals plots


Apart from exploratory band regression, other graphical displays, along
with numerical summaries, can be of help in investigating whether the
assumptions of classical normal linear regression are reasonably valid in
practice. These are scatter plots of the raw data along with the regression
line, on the one hand, and residual plots, on the other. To illustrate their
use in practice, let us use another concrete example with Tanzanian
data (in file TANZANIA) on the annual growth rates of, respectively,
recurrent expenditures (RE) and recurrent revenue (RR), both deflated
by the consumer price index, of the government budget in Tanzania
during the 1970s and 1980s. Our interest is to examine to what extent the
growth in recurrent expenditure was constrained by the growth in recur-
rent revenue. Before you read on, we suggest that you work out exercise
4.2 which is based on these data. This way, you can compare your analysis
with ours.

Exercise 4.2
Using the data file TANZANIA, compute the growth rates of govern-
ment recurrent expenditures and revenues and then answer the following
questions:
Data analysis and simple regression 129
1 regress RE on RR, computing the regression coefficients, their stan-
dard errors, the R 2 , the residuals, and the fitted values of RE;
2 graph RE against RR with the regression line;
3 graph the residuals against the predicted values of RE;
4 plot the residuals against time, t;
5 fit a three-band exploratory band regression;
6 check whether the residuals behave as a sample of a normal distribution.
Is this a satisfactory regression? Explain your answer.

The regression of RE on RR yields the following results (standard errors


in brackets):
RE= 0.0086 + 0.924 RR R 2 = 0.78 (4.42)
(0.1036) (0.114)
Before we can make any inferences from this sample regression, however,
we need to check that the data do not violate our model assumptions.
Let us do this step by step. First and foremost, let is see whether the data
support the linearity assumption. Figure 4.3 gives us the scatter plot of
RE on RR along with the estimated regression line. A scatter plot with
regression line is a powerful analytical graph which combines, at a glance,
the scatter of the data points with the fitted line. It allows us, therefore,
to see whether the regression line is reasonably supported by the scatter
as a whole without there being any systematic deviations of data points
away from the regression line. At first glance, our estimated regression
seems quite satisfactory in this sense. However, a closer look reveals that
there are relatively more points below than above the regression line.
Note also that there is one point at the left of the scatter which has a
relatively large positive residual; we may expect that this point will exert
a slight pull upwards on the regression line.
To enable us to take a closer look at the residual variation after fitting
the regression line we use a plot of residuals against predicted values.
This type of plot is akin to Sherlock Holmes's magnifying glass: it blows
up the residuals to enable us to detect abnormal behaviour. Remember
that the residuals have zero mean, hence, ideally most residuals will
be distributed linearly and evenly with roughly equal spread around
this zero mean. If a residual plot shows sorne curvature in the scatter of
the residuals around its mean zero, this indicates that the relation between
Y and X may not be linear. If the spread of residuals widens, or
narrows, with increasing values of the fitted Ys, this indicates the
presence of heteroscedasticity. Unusually large residuals will show up
because of their distant location (as compared with other residuals) from
the zero mean.
Figure 4.4 plots the residuals of the regression of RE against RR against
the predicted values of RE. The horizontal line in the plot shows the
130 Econometrics far developing countries

0.246037 o

o
o
o o

o
o
o

o o
o o

o o
-0.184229
-0.170612 0.222829
Growth in recurrent revenue
stata™

Figure 4.3 Scatter plot with regression line: RE on RR

0.159212 80

79

~
e:
o
88
~ 82 76
75
oV>
E 87 77
Ql
"O 90
·¡¡; 85 72
Ql
a:
81 78
84

74
-0.103522
-0.14916 0.214631
Fitted values of RE
srara•M

Figure 4.4 Residuals versus fitted vaiues


Data analysis and simple regression 131
position of the zero mean. For convenience, each data point is identified
by the year to which it refers: hence, the growth rate from 1979 to 1980
is indicated by the label 80. In this way, it is easy to identify unusually
large residuals, such as the year 1980. In this year, recurrent revenue fell
dramatically while recurrent expenditure still registered mild positive
growth. It appears as if the Tanzanian government was caught unaware
in 1980 when government revenue clearly collapsed with the onset of the
economic crisis of the early 1980s. Apart from this unusually large residual,
most other residuals are reasonably evenly spread around the zero mean.
Hence, the plot <loes not show strong evidence of heteroscedasticity. Note,
however, that the majority of points on the left-hand side of the scatter
slope slightly upwards towards the horizontal (zero mean) line, indicating
mild non-linearity.
To verify this last point, let us try out a simple three-band exploratory
band regression. Figure 4.5. shows the picture. To enhance this plot, we
drew the vertical and horizontal axes corresponding to zero growth rates
for each of the variables. The exploratory band regression shows a slight
non-linearity of the underlying relationship between RE and RR, indi-
cating that recurrent expenditure responded slightly more strongly to a
fall in recurrent revenue than to its positive growth. This slight non-
linearity is so mild, however, that it <loes not seriously invalidate the
linearity assumption.

0.246316 o

o
o

o
o
o

o o
o o

o o
-0.184229

-0.170624 0.222939
Growth rate recurren! revenues

Figure 4.5 Exploratory band regression: RE on RR


132 Econometrics far developing countries
0.2 -

0.15 -

0.1 -

0.05 -

o-r-~~~----t:=================================t~~~~

-0.05 -

-0.1 -

-0.15 -

sra1a 1
M

Figure 4.6 Box plot of residuals: RE on RR

To check whether the residuals behave approximately as a normal dis-


tribution, we use the methods outlined in section 3.4 of Chapter 3.
Figure 4.6 gives the box plot of the residuals of the regression of RE on
RR, while Table 4.2 gives the relevant summary statistics. The box plot
shows that the residuals are fairly symmetrical but with a somewhat longer
upper tail due to the large residual for 1980 which does not quite qualify
as an outlier. The median and mean are close together and, similarly,
Bowley's measure of skewness does not indicate the presence of skew-
ness as far as the middle 50 per cent of the residuals are concerned. The
pseudo standard deviation is somewhat larger than the standard devia-
tion, indicating that the tails of the distribution are heavier than those
from a normal distribution, but not particularly so. The computed value
of the skewness-kurtosis test statistic turns out to be 5.62 which is not
significant at the 5 per cent level. In other words, the hypothesis of normal
distribution of the error term E cannot be rejected at the 5 per cent level
of significance.
When dealing with time series, it is important to check for the pres-
ence of autocorrelation. We shall discuss how to do this in far more detail
in Part IV of this book. However, a plot of the residuals versus time is a
useful diagnostic tool to provide us with a rough check whether or not
the residuals are autocorrelated. When plotting residuals against time, it
Data analysis and simple regression 133
Table 4.2 Numerical summaries of the residuals
Order-based statistics Mean-based statistics
Median -0.0026 Mean -0.0000
IQR 0.0836
Pseudo-standard deviation (Sp) 0.0619 Standard deviation 0.0586
Skewness Bowley's measure (b 8 ) -0.1100 Skewness (a 3) 0.8500
Kurtosis (a 4 ) 4.1200

0.2

0.15

a: 0.1
a:
e
o
lJ.J 0.05
a:
o(/)
ca::::i o
"O
-¡¡;
Q)
a: -0.05

-0.1

-0.15
70 72 74 76 78 80 82 84 86 88 90

Figure 4. 7 Residuals against time

is useful (a) to draw a horizontal line indicating the zero mean, and (b)
to connect successive points with line segments. Figure 4.7 uses our
example to show you how to do this. If errors in successive years tend to
move in the same direction, this type of residual plot will show the pres-
ence of runs of positive or of negative residuals (respectively, strings of
points above or below the line ). If successive residuals are relatively uncor-
related, the curve of connected residuals will cross the line in a similar
way as fiipping a coin switches from a run of heads to a run of tails. If,
however, successive residuals are negatively correlated, a positive residual
will probably be followed by a negative residual and, hence, the curve
will be very jagged (crossing the zero line almost continuously). In our
134 Econometrics far developing countries
case, the curve connecting successive residuals shows a series of runs up
and down the zero mean, not unlike the type of runs you get by flipping
a coin. Hence, this quick check seems to indicate that the residuals show
little evidence of autocorrelation.
It appears, therefore, that in our example the assumptions of the clas-
sical normal linear simple regression model are reasonably satisfied. This
sets the stage for statistical inference based on the sample regression line.
Exercise 4.3 suggests sorne questions of statistical inference applied to this
simple example.

Exercise 4.3
Using the regression results listed in 4.40:
1 What economic interpretation would you give to the slope coefficient
of this regression?
2 What <loes its intercept tell you?
3 Formally test the hypothesis that the slope coefficient equals 1.
4 Formally test the hypothesis that the intercept term equals O.
5 Construct confidence intervals for (a) the conditional mean of RE, and
(b) the predicted value of RE for, respectively, RR = -0.10; RR =O.O,
and RR = 0.15.
6 Compute the hat statistics and the standard errors for each of the
residuals of the regression.

The slope coefficient measures the absolute change in the dependent


variable, the growth rate of recurrent expenditures, in response to a unit
change in the regressor, the growth rate of recurrent revenue. In this case,
the estimated slope coefficient tums out to be very clase to 1 and, in fact,
the hypothesis that this coefficient equals 1 is quite acceptable from
the sample regression as can readily be seen if you do the relevant
t-test (if you remain unsure how to perform the t-test look at Box 4.1).
Hence, a unit change in the growth rate of recurrent revenue leads to
a unit change in the growth rate of recurrent expenditures as well.
Furthermore, we note that the intercept is clase to zero. If, using a t-test,
you formally test the hypothesis that the intercept is zero, you will have
found that this hypothesis is also maintained. lt appears, therefore, that
the assumption that the regression line goes through the origin is quite
plausible in this case, meaning that the growth rate in recurrent expen-
ditures varies proportionally with the growth rate in recurrent revenue,
thus indicating a constant elasticity between both variables. But how do
we estimate a regression through the origin? This is the question we
tum to next.
Data analysis and simple regression 135

Box 4.1 Performing the t-test in regression analysis


The formula to use is:
t-stat = bz - Ho(/32)
se(b 2 )
Let us first test if the intercept is zero; i.e.:
H 0 : b 1 = O H 1: b 1 =I= O
When the null hypothesis value of the coefficient is zero the calcu-
lated t-statistic simplifies to the ratio of the estimate to its standard
error. So in this case:
0.0086
t-stat = - - = 0.63
0.0136
There are 20 observations, so there are 18 ( = 20 - 2) degrees of
freedom. The critica! value of the t-statistic at the 5 per cent level
is 2.101; i.e. if the null hypothesis is true then 95 per cent of all
sample estimates are expected to produce t-stats of between -2.101
and 2.101. The calculated value falls within this band (or, to put the
same thing another way, the absolute value of the calculated value
is less than the critical value ), so we accept the null hypothesis.

To test if the slope coefficient is significantly different from zero we


set up the hypotheses:
H 0 : b 2 = O H 1: b2 =I= O
and the corresponding t-statistic is
0.924
t-stat = - - = 8.11
0.114
The same critical value applies so in this case we reject the null
hypothesis; i.e. the slope coefficient is significant. When we say a
coefficient is 'significant', it is short-hand for 'significantly different
from zero at a given (most usually 5 per cent) level'.

Suppose now we wish to test if the slope coefficient is significantly


different from unity. In this case the null and alternate are:
H 0 : b 2 = 1 H 1: b 2 =I= 1
and so the test-statistic is:
0.924 - 1
t-stat = = -0.67
0.114
The critical value is again the same and we are unable to reject the
null; that is, the slope coefficient is insignificantly different from one.
136 Econometrics far developing countries
4.6 REGRESSION THROUGH THE ORIGIN
A regression with zero intercept, 13 1 = O, is called a regression through
the origin since it passes through the point (0,0). We often have a priori
reasons why a regression should go through the origin. For example,
when specifying the regression of RE on RR we might have had strong
reasons to believe that this regression will go through the origin. How-
ever, this should not lead us to try out a regression through the origin
befare farmally testing whether the hypothesis 13 1 = O is in fact data
admissible.
Why should we test first? The reason is that the restricted model
(without constant term) does not tell you how much worse or better it is
in comparison with the unrestricted model (with the intercept included).
To see whether the intercept can be dropped, we should first estimate the
regression with the intercept and subsequently test whether the constant
term can be dropped from the equation. This is a very rudimentary appli-
cation of the principle of general to specific modelling. Never jump to the
specific (restricted) model befare testing whether the restriction (in this
case, dropping the constant term) is in fact data admissible. Dropping a
constant term from an equation may appear trivial but, as we shall show
in Chapter 6, if inappropriate, it can lead to serious misspecification of
the underlying relation.
Without intercept, the relevant regression model becomes:
y= 132 X+ E (4.43)
with all the usual assumptions of the classical model about E and the
sampling process. The sample regression will obviously feature only one
coefficient b 2 , as b 1 = O. In this case, the least squares solution is as fallows:
b = LY;X;
2 LX.2 (4.44)
l

This regression through the origin, unlike the general model with slope
and intercept, does not adjust far the sample means. In other words, all
the farmulae far the relevant statistics in this case can be derived simply
by taking the corresponding farmulae far simple regression with a slope
and intercept and replacing all sample means by zero. The test statistics
and confidence intervals can also be derived in this way.
The R 2 statistic of a regression through the origin, however, loses much
of its usefulness as a measure of goodness of fit. It is not comparable with
the R 2 of the corresponding regression with intercept and slope. It is
furthermore possible to come across a negative R 2 in a regression through
the origin. This can occur when the intercept should not have been
dropped and the resulting residual sums of squares turns out to be higher
than the total sums of squares.
Our example yields the following regression through the origin:
Data analysis and simple regression 137
0.25 o

0.2
o
~
.a 0.15 o
'5
e: o
(])
a. 0.1
X
(]) o
e: 0.05
~
:::;
u o
~ o
.s::
~ -0.05
e
('.)
o o
o o
0.1
o
-0.15
o o

-0.15 -0.1 -0.05 o 0.05 0.1 0.15 0.2 0.25


Growth: recurren! revenue
s ra ra•M
Figure 4.8 Regression through the origin: RE on RR

RE= 0.934RR R 2 = 0.788 (4.45)


(0.111)

Exercise 4.4
Using the regression results in equation 4.45:
1 Formally test the hypothesis (3 2 = l.
2 What economic interpretation can you give to this unitary slope coef-
ficient? Why might we think it valid on theoretical grounds to exclude
the intercept from this regression?

4.7 OUTLIERS, LEVERAGE AND INFLUENCE


As in univariate analysis, you should always beware of points which do
not fit the general pattern or exert undue inftuence on the outcome of
our numerical summaries. In univariate analysis, we only had to deal with
the problem of outliers. In regression analysis, however, the problems we
can encounter with unruly data points are more complex. Consequently,
we need to sharpen our concepts to be able to spot the unusual and to
single out data points which exert undue pressure on our results. This is
particularly important when we rely on least squares regression since, as
we know, the least squares principle does not give us resistant nor robust
summaries. In regression analysis, there are three types of data points
138 Econometrics for developing countries
which should concern us. These are: an outlier, a point of high leverage,
and an inftuential point. It is important to get a good intuitive grasp of
these concepts in arder to be able to use them effectively in applied work.
Let us explain them one by one.
To start with, An outlier in a regression is a data point which has a
large residual. Large in this context does not refer to the absolute size of
a residual but to its size relative to most of the other residuals in the
regression. In Chapter 3, you saw how to identify outliers in a univariate
empirical distribution. The same techniques can be used to analyse resid-
uals of a regression. Note, however, an important distinction. When we
say a point is an outlier in univariate analysis, it is so defined with refer-
ence to its own mean (i.e. unconditional mean). When a point is an outlier
in bivariate analysis it has a large residual, i.e. Y value, far removed from
its fitted value (i.e. its conditional mean). But in general, a large residual
is easy to spot since it usually sticks out in a residual plot.
A point of high leverage is a different matter and its effect on a regres-
sion is more complex. It can be defined thus: 'A data point has a high
leverage if it is extreme in the X-direction; i.e. it is a disproportionate
distance away from the middle range of the X-values' (Myers, 1990: 250).
Note the difference between outliers and points of high leverage. In
regression, the concept of an outlier refers to a residual. That is, an outlier
is defined as an exceptionally large vertical distance of a data point from
the regression line. The definition of leverage does not involve the regres-
sion line at all. High leverage simply refers to a data point which is
disproportionately distant from the other data points in the X-direction:
that is, its X-coordinate deviates considerably from the X-coordinates of
the (majority of) the other data points. In other words, a point of (very)
high leverage in a regression of Y by X is an outlier in the univariate
distribution of X.
But why does leverage matter if its definition does not involve the
regression line at all? The reason is that points of high leverage can exert
undue inftuence on the outcome of a least squares regression line. That
is, points with high leverage are capable of exerting a strong pull on the
slope of the regression line. Whether they do so or not is another matter.
The point is that they have the potential to do so.
This brings us to the concept of an influential point. 'A data point is
inftuential if removing it from the sample would markedly change the
position of the least squares regression line' (Moore and McCabe, 1989:
185). Hence, influential data points pull the regression line in their direc-
tion. Note that influential data points do not necessarily produce large
residuals: that is, they are not always outliers as well, although they can
be. It is precisely because they draw the regression line toward themselves
that they may end up with small residuals. Conversely, an outlier is not
necessarily an influential point, particularly when it is a point with little
leverage (see exercise 4.10).
Data analysis and simple regression 139
The concepts of outlier, leverage and inftuence matter a great deal in
applied work. The reason is that the least squares regression line (like its
cousin, the sample mean) is neither resistant nor robust. In other words,
least squares performs well under ideal circumstances, but not when the
assumptions of classical normal linear regression are seriously violated in
practice. The presence of outliers or of inftuential points often gives us a
clear signal that our model is probably misspecified. For example, a cluster
of outliers or inftuential points may indicate that we wrongly applied our
model to a set of points which do not derive from the same (homoge-
neous) population. A single exceptional point may signal that a typing
error was made when recording its value or it may point us towards vari-
ables which we failed to account for in our model.

Exercise 4.5
Look carefully at the four plots in Figure 4.9 (overleaf). For each plot
write down whether any of the points is: an outlier, a point of high
leverage, an inftuential point, or sorne combination of these.

You will have noticed that each plot contains only one point which qual-
ifies for discussion. Obviously, with real data the situation can be more
complex. Table 4.3 shows our quick summary. In general we note:
1 outliers are not necessarily inftuential (Plot 4)
2 but they can be so (depending on leverage) (Plot 3)
3 yet high leverage points are not always inftuential (Plot 1)
4 and inftuential points are not necessarily outliers (Plot 2)
In terms of visual displays, outliers can best be spotted with residual
plots (but they are also visible in a scatter plot especially if the fitted line
is shown), while inftuential points require us to look at scatter plots since
they do not show themselves on residual plots if they do not have large
residuals. Just running regression without looking at any plots leads you
to ignore inftuential points and makes spotting outliers more tedious. This
is the simple but powerful lesson of this exercise.
As shown above, graphical displays can be of great help in spotting
outliers and inftuential points. Apart from these graphical methods,

Table 4.3 Plot summary


Plots Outlier High Leverage Influence
Plot 1 No Yes No
Plot 2 No Yes Yes
Plot 3 Yes Yes Yes
Plot 4 Yes No No
140 Econometrics for developing countries
110 o

>-
"O

~Q)
Vl
..e
o

12.3972
0.05911 20
Observed X

Plot 1

35 o

o
30 o o
o o
o
o o
o
25 o o
>- o o o o
"O o
Q) o
e:
Q)
20 o
o
Vl o
..e o o
o 15 o

10

o
5

o
o 5 10 15 20 25 30 35
Observed X
51 i3Ii3'R
Plot 2

Figure 4.9 Graph for exercise 4.5


Data analysis and simple regression 141

75.0795

o
o o o
ºo o o
o o 0o o o
Q> o o
0 0
o o 0
¡;p 0
o o
0

o
o 8
8' o °
o:9 ºo f' o
o o o
o o
º oº oioooº º
Q'.) o ~o
ocP ooº
o oº
ºo 8 8)ºº

'O o
12.3972 cP

0.05911 20
Observed X
5Ii3l i3"'

Plot 3

75.0795 % oo
Q:i o
o o
o o o
o o o
o o
o o o o
o ºo
Oo o
o o o
o o
o
>- ºº o
o o
"O 8
Q)
o 8º o
e:
Q) oºº o
o cP o
"'
.o o o o
o o
o o o 00 o o o
o
o o
ºº o o
o ºº
o oº o
ºº
o o o
o o
o o o
o 80
'b o
12.3972 ºº

0.05911 9.98287
Observad X
sra1a'"
Plot 4

Figure 4.9 (continued) Graph for exercise 4.5


142 Econometrics far developing countries
however, we can also rely on special statistics designed to detect outliers,
leverage and influence. Here we shall look at three such statistics: student-
ised residuals to detect outliers, the hat statistics to measure leverage, and
DFBETA statistics to measure influence.

Studentised residuals
In order to make the outliers conspicuous in relation to rest of the resid-
uals, it is useful to consider them in the context of the overall residual
variation. One way in which this can be done is to calculate the stan-
dardised residual, which is simply the residual divided by the standard
error of the estímate (i.e. standardised residual = e/s). However, the
problem with this measure is that if there is an outlier in the data set it
will ínflate the standard error of the regression. This problem is catered
for by using instead s(i) where the (i) subscript denotes a statistic calcu-
lated having dropped the ith observation from the sample. Be careful not
to be confused by this notation; for example h; is the hat statistic for
observation i (see below), whereas bz(i) is the slope coefficient having
dropped the ith observation from the sample.
Making this adjustment, we define the studentised residuals (t;):
e.
(. = ' (4.46)
' s(i)Y(l - h;)
where s(i) is the standard error estímate of the regression (defined in equa-
tion (4.26)) fitted after deleting the ith observation, and h; is a measure
of leverage as defined in equation (4.39). The additional term in the
numerator, '1(1 - h;), is necessary since the variance of the residuals is not
constant. With this adjustment, we get a t-statistic which tests whether the
ith residual is significantly different from O and, hence, signals an outlier
which <loes not really fit the overall pattern. (In fact the studentised
residual may be interpreted as a test of influence on intercept, but we
pursue this interpretation in Chapter 6.) It is possible to obtain formally
derived critica! values (which are larger than those from the usual t-table)
against which to compare the calculated value, but we recommend that
the studentised residual be used as an exploratory tool, as in the example
below.
Take another look at Figure 4.6 which depicts the box plot of the resid-
uals of the regression of RE on RR. This plot <loes not reveal the presence
of outliers, although its upper tail is somewhat more prolonged than its
lower tail. The box plot of the studentised residuals shown in Figure
4.10, however, tells a different story: the data point for the year 1980
now appears as an outlier. Studentised residuals, therefore, are much
better than the usual residuals at spotting outliers. The reason is that
each studentised residual is obtained by dividing the least squares residual
by its standard error (hence, a t-value) where the standard error of the
Data analysis and simple regression 143
1980

srata""'
Figure 4.10 Studentised residual RE

regression is estimated by deleting the ith data point. If the ith data point
is an outlier, the standard error of the regression after deleting the ith
observation will be significantly lower than the standard error of the
regression with all points included. This explains why the studentised
residual is better at bringing out outlying points.

The hat statistic


We already defined the statistic h¡ in equation 4.39 as follows:
1 (X¡ - X) 2
h¡ = --¡; + L(X¡ - X)z (4.47)

which serves as a measure of leverage of the ith data point. It measures


leverage because its numerator is the squared distance of the ith data
point from its mean in the X direction, while its denominator is a measure
of overall variability of the data points along the X-axis. The statistic h¡,
therefore, measures the distance of a data point in relation to the overall
variation in the X direction. Therefore, the higher the value of h¡ the
higher is the leverage of the ith data point.
Theoretically, as can be seen from equation (4.47), leverage can range
from lln to l. If the X value of a data point is exactly equal to the mean
144 Econometrics far developing countries
of X values, the point has no leverage and h¡ is equal to l!n. The statistic
h¡ approaches 1 as a point has very high leverage. Such a point will domi-
nate the scene completely and control the location of the regression line.
Obviously, this would be an extreme case. But when is leverage high
such that it can pose a problem? Huber (1981, see also Hamilton, 1992:
130-1) suggests the following guidelines based on the maximum observed
h; = max(h;):
max(h;) < 0.2 little to worry about
0.2 < max(h;) < 0.5 risky
0.5 < max(h;) too much leverage
Figure 4.11 graphs the hat statistics against the values of the regressor
in our regression of RE on RR. The graph clearly shows that leverage
increases as the X values are more distant from their mean. Hence,
max(h;) is slightly above 0.20, but it corresponds to a point which presents
little danger in our regression. As such, there is little to worry about as
far as leverage is concerned.
Why do we refer to the h; values as hat statistics? To see this point,
recall that the least squares estimators are linear functions of the observed
Y values. Consequently, the predicted values of Y, called Y-hat, are also

75
0.2 89

80 78

0.15
84

82

88
0.1
7~1
85 90
73

0.05

o
-0.02 -0.15 -0.1 -0.05 o 0.05 0.1 0.15 0.2
Growth rate: recurren! revenue
STaTa™
Figure 4.11 Hat statistics versus RR
Data analysis and simple regression 145
linear functions of the observed Y values. The hat statistic, h; is nothing
other than the coefficient of the observed Y; in the linear equation which
expresses the predicted value Y; as a linear function of all observed Y; s.
This explains why they are called hat statistics.

DFBETA statistics
The DFBETA statistic is defined as:
DFBETA. = b 2 - b 2<0 (4.48)
' se(b 2 )(i)
where b2(i) and se(b 2)(i) are the slope coefficient and standard error of the
estimate of the slope from regression estimated, having dropped the ith
data point from the sample. The DFBETAs measure the sensitivity of the
slope coefficient to the deletion of the ith data point. If the deletion of
this point leads to a drastic change in the slope coefficient, it follows that
the ith data point is infiuential. In other words, a large value of the
DFBETA statistic for a given data point indicates that this observation
has a sizeable impact on the slope coefficient of the regression. DFBETA
may also be calculated for the intercept, although this is not a common
practice. (In Figure 4.9, plot 4, the outlier will infiuence the intercept,
though not the slope coefficient.)
But how big does a DFBETA statistic have to be to be considered
large? The DFBETA is not a formal test statistic like, say, the t-test.
Therefore we do not have critical values derived from statistical theory.
But we can use sorne rules of thumb. In general, if DFBETA > 2, the
corresponding data point is unquestionably an influential point. This is a
general criterion. It is also possible to relate the cut-off value of the
DFBETAs to the sample size, n. If DFBETA > 21-..Jn, the corresponding
data point may be deemed infiuential (Myers, 1990: 261; though sorne
sources suggest 31-..Jn). As a rule, it is useful to make a box plot of the
DFBETA statistics and check whether there are any outliers. Figure 4.12
shows the box plot of the DFBETA statistics of the regression of RE
against RR. As we can see, none of the data points has a DFBETA statistic
which exceeds 2, while only the data point 80 (corresponding to the growth
rate from 1979 to 1980) exceeds 21-..J20 = 0.447. This confirms our earlier
hunch that this point exerted a slight pull on the regression line without,
however, causing any major distortion of the results.
A final point: DFBETA statistics are obtained by deleting each data
point in turn and checking the effect of doing so on the slope coefficient of
the regression line. It is possible, however, that a number of infiuential
points may cluster together and jointly pull the regression line in their
direction. In such a case, deleting data points one by one may not reveal
the pull exerted by this infiuential cluster of points. This is why you should
never rely solely on DFBETA statistics to check for infiuence, but also take
146 Econometrics for developing countries
0.1

0.5 74

78
-0.5

-1

-1.5 80

-2
5 laTC:J'M

Figure 4.12 Box plot of DFBETA statistics

a good look at the scatter plot of Y against X to see whether there are any
clusters of influential points. The formula for DFBETA may be just as eas-
ily applied by dropping two or three points from the regression. At times,
several clusters exist which may pull the regression line in similar or oppo-
site directions. But if you find you have many 'influential points' you can
be sure that the problem is one of model misspecification.

Exercise 4.6
In section 3.3 of Chapter 3 we investigated the relation between the differ-
ence between female and male life expectancy, L, of developing countries,
on the one hand, and GNP per capita, Y, on the other. In doing so, we
grouped the data into three income categories: low, lower-middle and
upper-middle income countries. Here we look at the same relation without
prior grouping of countries in income categories. Using the data file
SOCECON:
1 regress L on Y;
2 check whether this regression is likely to satisfy the model assump-
tion;
3 try an exploratory band regression;
4 check graphically whether there are outliers, points of high leverage,
or influential points;
Table 4.4 Summary measures of outliers~ influence and leverage
Statistic Formula Use Critica[ value

Studentised residual (t¡)


f¡ = s(i)Y(l - h¡)
Outliers Critica} values available (higher than usual
t-test), but recommend use studentised
residual as an exploratory tool

1 (X-X)- 2
Hat statistic ( h) h=-+ 1 - Leverage Bounded by lln (no leverage) and 1
n L.(Xi -X)2
(extreme leverage ); values above 0.5
indicate excessive leverage and values over
0.2 indicate that the observation may give
problems.
b2 - b2(i)
DFBETA DFBETA¡ = se(b2){í) Influence Under 2/-Jn the point has no influence;
over 3/-Jn the point is influential and
strongly so if DFBETA exceeds 2.
Note: n is the sample size; k is the number of regressors; the subscript (i) (i.e. with parentheses) indicates an estímate from the sample
observation í. In each case you should use the absolute value of the calculated statistic.
148 Econometrics far developing countries
5 compute the studentised residuals, the hat statistics, and the DFBETAs
for the data points. Is this regression a good summary of the data?
Note that the computation of studentised residuals, hat statistics and
DFBETAs is cumbersome unless you have access to a statistical pack-
age which routinely provides these diagnostic statistics. Unfortunately,
most econometric packages do not provide these statistics. For this
reason, it is useful to familiarise yourself with a statistical package (for
example, STATA) which incorporates residual analysis and influence
diagnostics in its statistical routines.

4.8 TRANSFORMATION TOWARDS LINEARITY


In Chapter 3 we saw that simple data transformations help us to extend
the reach of the classical model to make inferences around the popula-
tion mean of a normal distribution. There we used transformations to
alter the shape of the distribution of our observed variables. The main
purpose, therefore, was to transform our data towards normality. In a
similar fashion, data transformation can be used to extend the reach of
regression analysis. Here, however, our aim is to transform the bivariate
data towards linearity. Indeed, ordinary least squares is based on the
assumption that the regression line is linear. But at times data transfor-
mations allow us to extend the reach of simple linear regression to handle
a wide variety of curvilinear relationships between Y and X.
In fact, there is an important principle here in statistical analysis which
is often overlooked. In statistics, we generally start with simple models
which perform well if a rather restrictive set of ideal conditions (usually
involving the normality assumption) are satisfied. Obviously, there are
many situations in which this restrictive model will not be appropriate.
But this <loes not mean that we immediately discard the model. Instead,
the basic idea is to get as much mileage as possible out of these simple
but powerful models. In Chapter 3 we showed how data transformation
towards normality extends the reach of the classical normal model for
making inferences about the population mean. In this section we show
how to extend simple regression analysis through transformations towards
linearity.
If we transform either X or Y, or both, to achieve linearity of the regres-
sion line, we should not forget that, in the process, we also transformed
the distributions of Y or X. Hence, while here our main aim is to achieve
linearity, we should not set aside the lessons we learned in Chapter 3 as
to how data transformations alter the shape of the distributions of the
variables. In practice, a useful rule of thumb is to keep in mind that linear
regression tends to work best when both variables are similarly (prefer-
ably symmetrically) shaped (Hamilton, 1992: 148). The reason this is so
is straightforward. If, for example, X is skewed while Y is more or less
normally distributed, it is hard to expect that the regression of Y on X
Data analysis and simple regression 149
will leave an error component which is normally distributed. However,
the converse is not true. Y and X may be associated with one another
and similarly shaped, but this does not mean that the regression between
them will be linear. To see this point, take another look at Figures 4.1
and 4.2, which depict the relation between the recruitment of and the
demand for casual labour on the day-shift in Maputo harbour in the early
1980s. Both variables are approximately normally distributed, but the
regression line is non-linear as well as heteroscedastic.
In econometric practice, the logarithmic transformation is very popular.
One reason is that functions which can be linearised with the aid of loga-
rithms have coefficients which lend themselves to meaningful economic
interpretations, such as an elasticity or a growth rate. Another reason is
that the logarithmic transformation frequently, but not always, does the
trick with socioeconomic data which are often skewed to the right. In this
section, therefore, we shall discuss severa! functional forms which can
easily be rendered linear by using the logarithmic transformation.

The double-Iog transformation


A commonly used function in mathematical economics takes the following
form:
Y= A X132 (4.49)
This is a function with a constant elasticity given by í3 2• That is:
dYX
dX y= /32 (4.50)

Taking logarithms of both sides of equation (4.49) yields:


lnY = í3 1 + í3 2 lnX (4.51)
where
í3i = lnA (4.52)
As we can see, this double-log equation is linear with respect to its
transformed variables. Linear regression, therefore, is feasible, but what
assumptions do we make about the error term?
The normal procedure is to add an error term to the double-log spec-
ification and to assume that it satisfies the usual assumptions of the linear
regression model. Hence, we get:
lnY = í3 1 + í3 2 lnX + E (4.53)
Taking anti-logarithms of both sides yields:
y= AXl32E' (4.54)
where
E'= e• (4.55)
150 Econometrics for developing countries
which shows that the error term in the original model is multiplicative
rather than additive in nature. If we assume that the error term in the
double-log specification is normally distributed, it follows that the error
term in the original model will have a lognormal distribution (see Box
4.2). An error term which is lognormal in nature and which enters the
specification in a multiplicative fashion is bound to affect the variation in
the dependent variable due to the behaviour of outlying values in its tail.
If, instead, we assume that the error term in the original equation is addi-
tive rather than multiplicative, the resulting model specification becomes:
(4.56)
which can no longer be linearised by taking logarithms of both sides of
the equation and, hence, we need to apply non-linear regression, a topic
which is beyond the scope of this book.

Box 4.2 A lognormal variable


A variable has a lognormal distribution if the logarithm of the vari-
able is distributed normally. The lognormal distribution is skewed
to the right and takes only positive values. It rises rather steeply to
a maximum and subsequently declines smoothly into a long right
tail which grows thinner and thinner as it reaches towards higher
and higher values.

But, in fact, the assumption of a multiplicative error term in specifica-


tion (4.54) is quite appropriate in many practica! application, as we shall
now illustrate with a simple example. The top panel in Figure 4.13 depicts
the scatter plot of energy consumption per capita against GNP per capita
for low, lower-middle and upper-middle countries for which data are in
the file SOCECON. We enhanced the scatter plot with simple box plots
for each variable (the horizontal box plot depicts the shape of the
regressor, GNP/capita, and the vertical box plot that of the regressand,
energy consumption per capita). As you can see, both variables are fairly
strongly skewed to the right. In fact, most of the data points are huddled
together in the bottom-left comer of the scatter plot. Moreover, the scatter
clearly is heteroscedastic. A careful look at the plot also reveals that the
scatter of points bends upwards from left to right. A linear regression,
therefore, does not appear to be appropriate.
The bottom panel of Figure 4.13 shows what happens if we plot the log-
arithms of both variables against each other. The double-log transforma-
tion worked remarkably well. The scatter of points is now homoscedastic
and more evenly spread out. What matters even more is that the data points
trace a reasonable linear relation between both transformed variables. As
1---{ 1
5000
~ o o
·a_
ctS
o 4000
O>
a..
e: o
o 3000
li. o o
o o
E o o
::J
en 2000 o
e: o
oo
o o o
o

ó
>. o o
e 1000 ºo o o
o
(])
o ºo i o o oº
~o¡¡g~@o
e: o @
w <o
o
f
o 1000 2000 3000 4000 5000 6000 7000
GNP per caprta

1
§' 9
·a_
ctS o o
()
8 o o
O> ºoº o0 o
a.. o
aeº o
e: 7 o
o 'b go
o o
o o o o 'b o oº
li. 6 o o o
E
::J
00
o
o o o o 0
o
0
o
o
io
0

e: 5 o o
o o
oo o º º oº o
o o
>. 4 o o o o
Ol o o o
w
e: 3
o o
o ~ Oa o
o
o
!:!::!. o
Ol
_Q
2
1

4 5 6 7 8 9
lag (GNP/capita} STaTa""'

Figure 4.13 Energy consumption versus GNP: per capita


152 Econometrics for developing countries
we can see, the double-log transformation appears to have solved three
problems at once: non-linearity, heteroscedasticity and unequal density of
the original scatter. In Chapter 7, on heteroscedasticity, we shall return to
this property of power transformations to render residuals homoscedastic
in cases where one or both of the variables is skewed. Note that, after trans-
formation, both variables have become reasonably symmetric in shape. But
our main point here is that the double-log scatter now is more linear in
shape and, hence, allows us to extend the reach of linear regression to a
curvilinear relationship.

Exercise 4.7
Use the data on energy consumption per capita, E, and GNP per capita,
Y, from the data file SOCECON far all countries with a GNP per capita
of less than $10,000. These are the low, lower-middle and upper-middle
income countries. Answer the following questions:
1 Regress E on Y, and log(E) on log(Y).
2 In each case, check whether the normality assumption is reasonably
satisfied, whether the residuals are homoscedastic, and whether there
are any outliers or influential points.
3 Which, do you think, is the better regression, and why?

Perhaps you were inclined to base your judgement on a comparison of


the R 2s of both regressions? If so, keep in mind that such a comparison
is invalid since the dependent variable is not the same in both cases: the
linear specification features energy consumption per capita as the depen-
dent variable while the double-log specification has the logarithm of
energy consumption per capita as regressand. To be able to compare two
R 2 s, we have to compare like with like: both should be expressed as a
share of the total variation in the same dependent variable. If the depen-
dent variables differ across specifications, so will their total sums of
squares and, hence, comparisons between shares of different things will
be meaningless.
But, more importantly, to choose between specifications it is far more
important to look carefully to see which specification best satisfies the
assumptions of the regression model rather than mechanically picking out
the one which features the highest R 2 . In this case, the double-log regres-
sion is more in line with the assumptions of the classical regression model.
Note that the linear fit with the transformed data yields a non-linear rela-
tionship if we return to the original raw data. Figure 4.14 compares the
simple regression of energy consumption per capita on GNP per capita
with the non-linear fit obtained from the double-log regression. While not
immediately obvious from the graph, the non-linear curve fits the large
number of data points huddled in the bottom-left comer far better than
the linear specification. The reason is that the double-log curve gives
Data analysis and simple regression 153
5000

.l!l
4000
·a.
C1l
u

c.
e 3000
o
aE
::J
(/)
e 2000
o
u
>,
e>
Q)
e
w
1000

o
o 1000 2000 3000 4000 5000 6000 7000
GNP per capita
srara"'
Figure 4.14 E on Y versus log(E) on log(Y)

roughly equal weight to all data points, unlike the linear regression, the
location of which is mainly determined by the inftuential points on the
right-hand side of the scatter.
More formally, there is a procedure which may be followed to compare
the R 2s for regressions with transformed dependent variables. The method
is as follows:
1 Carry out the regression with the transformed dependent variable and
calculate the fitted values.
2 Convert these fitted values back to the original data units (for example,
if you have made a log transformation, then take the exponential of
the fitted values).
3 Calculate the correlation coefficient between the converted fitted
values from step 2 and the actual values of the dependent variable.
The square of this correlation coefficient may be directly compared
with the R 2 from the regression with the untransformed dependent
variable.

Semi-logarithmic transformations
Semi-logarithmic transformations involve equations which can be
linearised by a logarithmic transformation of either the dependent or the
154 Econometrics for developing countries
independent variable. There are two variants. The first is given by:
(4.57)
This specification, like the double-log specification, involves a multiplica-
tive error term. Taking logarithms of both sides yields:
lnY = 13 1 + 13 2 X+ E (4.58)
where
131 = lnA (4.59)
The slope coefficient of this specification can be expressed as follows:

-dln
f3z- -- Y -dY
--1
(4.60)
dX -YdX

which shows that the slope coefficient depicts the relative change in Y per
unit change in X.
To illustrate this semi-log model, let us take another example with data
from the file SOCECON. The top panel of Figure 4.15 shows the scatter
plot of energy consumption per capita, E, against the urban population
as a percentage of the total population, U, for low, lower-middle and
upper-middle income countries. As you can see, the scatter plot reveals
that the underlying relationship is non-linear. Note, furthermore, that the
distribution of the regressor, U, is fairly symmetric in shape, while that
of the regressand, as we know already, is skewed to the right. The lower
panel of Figure 4.15 plots log(E) against U. Once more, the transforma-
tion solved more than one problem: the scatter shows a linear pattern and
is no longer heteroscedastic. Consequently, the regression of log(E)
against U, unlike that of E against U, is likely to satisfy the assumptions
of the classical linear model. We leave it to you as an exercise to verify
this. Figure 4.16 compares the linear regression line with the non-linear
curve estimated by the semi-log model.
An interesting case of this semi-log model arises if its regressor is a
variable denoting time, t. For example, if t denotes continuous time the
semi-log model depicts an exponential trend with a constant (instanta-
neous) rate of growth given by its slope coefficient.
It is more common, however, to measure time in discrete intervals (say,
a year). In this case, we can best modify the specification as follows:

(4.61)

where r is the constant (yearly) growth rate implied by the trend. Taking
logarithms of both sides of the equation yields:
(4.62)
where
s 15000 j o
·a. 14000
cu 13000
u 12000
ID 11000
a.. 10000
e
o 9000
li. 8000--, oº
E 7000
~
ti) 6000 o
e 5000 o ºo
o o o Cb
u 4000 o oº o o
>. o o o o
3000 o o ºº oo o ºo
e>
Q) 2000
o
o o o ºº o
e o 0 o
1000 o o
w ºo º oººo
ºoO oO o
o o oco ~ ~ 8 o~ c2 ~ c&ne9 oO oº o

o 10 20 30 40 50 60 70 80 90 100
Urban population in % total

§' 10
·a.
cu o
u 9 ºº
o
ID o o o ºº
o g:i ºo <-O o
c.. 8 o o o cP q, o 0
o
e o o o oO o
o o o
7 o ºº o o o o
li. o o o
o ºª o o o
o
E o
~
ti)
e
6
o
o oº o
o o oo
o
o o o
o
ª oº o
o 5 o oººo o
u o o oO o
00 o o o o
>.. o o oºº o
Ol 4 o o 00 00 @
ID
e o o o o
3 o o o o o o
~ o
o
Ol 2
_Q
1 ¡ 1 l 1 1

o 10 20 30 40 50 60 70 80 90 100
Urban population as % total STciTa.,.,..,

Figure 4.15 Energy consumption versus urban population as percentage of total population
156 Econometrics far developing countries

15000
14000
13000
12000
.E 11000
·a.
Cll 10000
~
e 9000
o
E. 8000
E
:J 7000
"'oe 6000
u
>- 5000
e>
Ql
e 4000
UJ
3000
2000
1000
o

o 10 20 30 40 50 60 70 80 90 100
Urban population as % total
srata"'
Figure 4.16 E on U versus log(E) on U

13 1 = lnY0 13 2 = ln (1 + r) (4.63)
In summary, the slope coefficient obtained by regressing log(Y) against
t gives us an estimate of either the instantaneous rate of growth or, after
taking the anti-logarithm and subtracting 1, the period rate of growth,
depending on whether time is seen as a continuous or a discrete variable.
Note however that ln(l + r) = r for small r so that the growth rate will
be approximately the same by either interpretation under these circum-
stances.
The second variant of the semi-logarithmic model only involves the log
transformation of the regressor, as follows:
(4.64)
Note that this model <loes not involve a transformation of the depen-
dent variable and hence the error term enters the equation in its usual
additive fashion. The slope coefficient of the model is given by:
dY dY
f3 - - --- -
2 - dlnX- dX/X (4.65)
which, less formally, depicts the absolute change in Y per unit relative
change in X.
.----¡

80
o
o o
o o o o o o o o
rP 8 o
70 -1 o ºº oº
0(11) 00 <t:i o o o o
{)' o
e o o ºo o~ o oº o

~<D 60 -1 o oº
o
a. &, o og
X o
<D 50 -1 Jo%o o
.!!:! o o
:.:::¡ o
o o
40

30

o 1000 2000 3000 4000 5000 6000 7000


GN P per capita

80
o o
o o
o o o o 8 ºo o oºo o
70
>. ºo roºº%º oc o
o o o
e O CD O oo'tJ o O
ctl
t3 60 o
e o o o
(l)
a.
X
o oo
o o o
o~ o o
o o
<D 50 00 080 ooººº
<D o o o
:s 40
o
o
o

30
1
4 5 6 7 8 9
log {GNP/capita) STaTa"™

Figure 4.17 Life expectancy against GNP per capita


158 Econometrics far developing countries
Let us consider an example which depicts a regression of Y against
log(X). The example, as before, is drawn from the file SOCECON to
enable you to follow it up as an exercise. The top panel of Figure 4.17
depicts the scatter plot of life expectancy, L, against GNP per capita, Y,
a simple plot of the relation between the health and the wealth of nations.
In this case, the data clearly are curvilinear in shape. The scatter of
L against log(Y) shown in the bottom panel of Figure 4.17 goes a long
way towards unbending the curve in the raw data. It appears, therefore,
that in this case the semi-log model, L against log(Y), is more in line
with the assumptions of classical linear regression. Figure 4.18 compares
the linear regression with the non-linear curve produced by the semi-log
model.
This particular functional form, which features the raw data as depen-
dent variable and the log transformed data as regressor, is appropriate in
the analysis of Engel curves with household budget data. Engel's law
postulates that 'the proportion of total expenditures that is devoted to
food tends to decrease exactly in arithmetic progression as total expen-
diture increases in geometric progression' (Working, 1943, quoted in
Barten, 1985: 462). In the regression of Y on log(X), the slope coefficient
measures the absolute change (arithmetic progression) of Y against
relative chance (geometric progression) of X. This explains why this
specification is so commonly used to estimate Engel curves. In fact, this

80

70

>.
()
e:
60
"'c.
t5
Q)

X
Q)

~
:.:J 50

o
o
40

30
o 1000 2000 3000 4000 5000 6000 7000
GNP per capita
srata™
Figure 4.18 L on Y versus L on log(Y)
Data analysis and simple regression 159
specification would have been most appropriate to model the data in Table
4.1 on food expenditures as a percentage of total household expenditures
as a function of total household income in Tanzania in 1969.

Exercise 4.8
Do exercise 4.6 again, regressing L (the difference between female and
male life expectancy) on Log(Y) rather than on Y (GNP per capita). Are
there any outliers or (clusters of) influential points left. If so, how do you
suggest tackling the problem?

In conclusion, data transformations can help us to get more mileage out of


the simple linear regression model since they make it possible to linearise
curvilinear relationships. The logarithmic transformation is particularly
popular on this count. They often do the trick and, furthermore, the slope
coefficients of models involving log transformations lend themselves to
easy, straightforward interpretations such as an elasticity ora growth rate.
Other power transformations can also be used. Regressing Y on l!X, for
example, allows Y to grow or decline towards a limiting value given by the
intercept of the model. A regression of Y on X2 depicts a simple quadratic
relation which can come in handy when X is skewed to the left while Y
is reasonably symmetric. Power transformations, therefore, extend the
reach of the regression model. Moreover, as we shall discuss in Chapter 7,
power transformations often allow us to eliminate heteroscedasticity of the
residuals, as illustrated by our examples in this section.

4.9 SUMMARY OF MAIN POINTS


1 The simple regression of Y on X is the line or curve of conditional
means of Y for given values of X. Simple linear regression assumes
that the conditional mean of Y is a linear function of X.
2 The most commonly used method to derive the sample regression is
based on the least squares principie. Given the assumption of classical
linear regression, the least squares estimators of the coefficients of the
regression line are BLUE. If, furthermore, the error term is distrib-
uted normally, the least squares estimators will be minimum variance
estimators.
3 The least squares regression line splits the total variation of the depen-
dent variable, Y, into two components: the explained variation due to
the variation in X and the residual variation. The R 2, a measure of
goodness of fit, is the proportion of the explained variation in the total
variation of Y.
4 Inference in regression analysis is based on the assumptions of the
classical normal linear regression model - in particular, on the
normality assumption of the error terms. The treatment of estimation,
160 Econometrics far developing countries
confidence intervals and hypothesis testing is very similar to that of
the classical model of estimating a population mean.
5 The least squares regression line, like the sample mean, is a powerful
tool of analysis if the assumptions of classical normal linear regression
are reasonably satisfied in practice. If not, the least squares regression
line rapidly loses its superiority. For this reason, it is important to
check whether the assumptions of the model are valid in practice.
Regression graphics - exploratory band regression, scatter plot with
regression line and residual plots - are important tools to check the
approximate validity of the assumptions.
6 A regression through the origin involves dropping the intercept from
the model. Always test first whether the data admit this type of model
restriction and keep in mind that the R 2s of the models with and
without the intercept are not comparable.
7 Outliers, leverage and inftuential points can lead us astray when
making inferences from a sample regression about the population.
Outliers are data points with an exceptionally large residual, while
inftuential points pull the regression line in their direction. Leverage
- a property of the X dimension - indicates whether a data point lies
nearer the edges of the scatter or in the middle as far as the X direc-
tion is concerned.
8 Hat statistics measure leverage. Studentised residuals and DFBETAs
help us to detect the presence of, respectively, outliers and inftuential
points. The latter two are deletion statistics inasmuch as their calcu-
lation involves deleting each data point in turn from the sample. (These
measures are summarised in Table 4.4.) DFBETAs should always be
used in conjunction with diagnostic regression graphics. It is always
possible that a cluster of points is exerting inftuence rather than a
single data point.
9 Data transformations can help to achieve linearity. In this respect, the
logarithmic transformation is most frequently used. Three basic spec-
ifications are possible depending on whether we transform Y or X or
both. Always check how the data transformation affects the shape of
the distributions of the variables concerned.

ADDITIONAL EXERCISES

Exercise 4.9
Use the data in data files INDONA and SRINA to estimate a consump-
tion function for Indonesia and Sri Lanka respectively. Repeat the
estimation excluding the intercept in each case, and comment on your
findings.
Data analysis and simple regression 161
Exercise 4.10
Demonstrate algebraically that adding a point (Xn + 1, Yn + 1) to a sample
of n observations will (a) not influence the slope coefficient of the regres-
sion of Y on X if Xn + 1 is equal to the sample mean for X of the n
observations; (b) that the intercept from the regression will probably
change even if Yn + 1 is equal to the sample mean for Y of the n obser-
vations; and (e) that if Xn + 1, Yn + 1 lies at the point of means of the sample
of observations, then the regression line is unchanged. Generate a numer-
ical example to illustrate your findings.

Exercise 4.11
Use the data in the data file TOT to regress the terms of trade on a
constant and a time trend where the trend is defined as (a) t = 1, 2, ...
39; and (b) t = 1950, 1951 ... 1988. Compare the estimated coefficients.
Derive algebraically the general result which is verified by your terms-of-
trade regression.

Exercise 4.12
Using the data in the data file TOT, regress the terms-of-trade index and
its log on a time trend. How do you interpret these two sets of results,
and which regression model has the more appropriate specification?

Exercise 4.13
Prove that the least squares estímate of the slope coefficient in the simple
regression model is an unbiased estimator. State clearly any assumptions
you make.

Exercise 4.14
Draw a scatter plot of consumption against income using the data in data
file SRINA. Plot on this graph (a) the fitted values of consumption
from the consumption function; and (b) the upper and lower limits of the
confidence interval for the fitted values. Comment on the shape of each
of the curves you have plotted.

Exercise 4.15
Figure 4.19 shows a data set with a clear point of high leverage. Also
shown are the regression line with and without this observation included
in the sample. When the observation is excluded, the regression line seems
to fit nicely through the points. However, for the full sample the regression
162 Econometrics far developing countries
240 .--------,--~~~~~~~~~~~~~~~~~~~~~~-----,

220 D

200

180

160

140

>- 120

100

80

60

40

20
o ~~~~~~~~~~~~~~~~~~~~~~~~~

o 20 40 60 80 100 120 140 160 180 200


X

Figure 4.19 Graph for exercise 4.15

line not only misses the point of high leverage but seems also to fit less
well to the other points. However, the R 2 from the full sample regression
is 0.85, compared to 0.78 when the point of high leverage is excluded.
How can this result be explained and what important lessons can you
draw from this example?
5 Partial regression: interpreting
multiple regression coefficients

5.1 INTRODUCTION

This chapter and Chapter 6 extend our analysis of regression to models


with more than one explanatory variable. This is a majar advance since
in applied work we usually are concerned with relations which involve
several explanatory variables. The good news is that not much additional
statistical theory is required when moving from simple to multiple regres-
sion. First, the least square line is derived in much the same way as we
did in simple regression. Second, the Gauss-Markov theorem extends to
multiple regression with minar modifications in the assumptions, and, simi-
larly, the least squares estimators turn out to be maximum likelihood
estimators and hence have minimum variances if the assumption that the
error terms are distributed normally is valid. Third, statistical inference
in multiple regression is a natural extension of earlier theory. At first
glance, therefore, multiple regression poses few new problems. But this
is not wholly true. In applied work, multiple regression poses new chal-
lenges and raises new problems. The main reason is that we now are
dealing with several explanatory variables at once.
But this raises the thorny question of variable selection - which vari-
ables to include in a regression? This is where modelling strategies come
into full play. With non-experimental data (that is, when we do not control
the explanatory variables), regression results differ, often markedly,
depending on the number of regressors we choose to include in the model.
In other words, the regression coefficient of an explanatory variable X
will, in general, differ depending on whether we regress Y, the dependent
variable, on X alone, on X and W, on X, W and Z, and so on. Theory
often offers a range of competing explanations which leave us still with
a great <leal of specification uncertainty as to which variables really matter.
Furthermore, theoretical arguments mostly involve postulating relations
between two or more variables, assuming other things to be equal. In
empirical research, we cannot make such an assumption because we
cannot hold other things constant. Our problem, therefore, is to disen-
tangle the separate effect of each explanatory variable on the dependent
164 Econometrics for developing countries
variable in a context where we are not quite sure which variables to
include and where we cannot hold any of them constant.
Traditional textbooks offer little assistance on this count in as much as
they generally assume that the model is correctly specified and, subse-
quently, go on to show that, under the by now familiar assumptions, least
squares estimators will be BLUE as well as mínimum variance estima-
tors. These texts also teach us what happens to the properties of our
estimators if we omit a relevant variable from the model or if we include
superfluous variables, but generally have little else to say about data
analysis in the context of specification uncertainty. But in applied work
with non-experimental data, variable selection - finding out which vari-
ables to include - is more often than not the name of the game. To come
to grips with this issue is our main concern in this chapter and the next.
This does not mean, however, that we shall not discuss the multiple
regression model, its assumptions and its properties. On the contrary, the
question of specification uncertainty arises precisely because of our aware-
ness that models are as good as the assumptions upan which they are
based. Misspecification matters because it leads us to make shaky or
invalid inferences from our data. Hence, a good understanding of the
statistical foundations of the models we use to approach data is essential.
Our concern here, however, is with understanding the assumptions and
the properties which follow from them, not with formal proofs and theo-
rems. Our aim is to focus on modelling strategies in applied work. It is
this emphasis which gives this chapter and the next their particular flavour
as well as structure.
This chapter is structured as follows. In section 5.2 we use an example
to show how we can move from simple to multiple regression through
partial regression, which involves a hierarchical sequence of simple regres-
sions. Partial regression shows that a multiple regression coefficient for a
particular regressor, X, is arrived at by sweeping out the linear influence
of the other regressors included in the model on both the dependent vari-
able, Y, and the particular regressor in question. In this example we
investigate the effect of changes in the price of basic foodstuffs on the
demand for manufacturing goods in India. Section 5.3 deals with the
application of the least squares principle to derive the sample multiple
regression line and discusses its mathematical properties. Section 5.4
returns to the concept of partial regression to show its formal equivalence
with multiple regression and discusses the coefficient of partial correla-
tion. The main purpose of this section, however, is to introduce you to
EDA's partial regression plot (also called the added variable plot), a
powerful graphical diagnostic tool which allows us to look carefully at a
bivariate scatter plot for any slope coefficient in a multiple regression.
Section 5.5 discusses the classical multiple linear regression model,
its assumptions and its properties. Section 5.6 reviews the use of the t-test
in multiple regression and shows how it can be put to work, through
Partial regression 165
re-parameterisation of the model specification, to test more complex
hypotheses which involve linear combinations of regression coefficients.
Subsequently, section 5.7 introduces you to a simple variant of fragility
analysis in multiple regression. The basic idea is to find the bounds within
which a particular regression coefficient varies across a range of plausible
rival neighbouring specifications (which differ with respect to the choice
of variables included in the model) so as to judge the robustness of co-
efficient estimates to changes in model specification. As usual, the final
section summarises the main points of this chapter.
This chapter, therefore, concentrates on the interpretation of regression
coefficients in the context of rival model specifications. As to modelling,
it teaches you how multiple regression seeks to disentangle the separate
effect of a particular regressor on the dependent variable when other
regressors are also included in the model. It shows how you can look care-
fully at how well the data support an individual regression coefficient in
a multiple regression. And, finally, it teaches you how to check the fragility
of a regression coefficient to neighbouring plausible alternative specifi-
cations of the model. In the next chapter, we turn to 'general to specific
modelling' in multiple regression as an alternative data-based approach
to modelling which involves model selection through testing down to see
which variables to include. There we shall make use of the F-test as an
important analytical tool to allow us to test linear restrictions on the coef-
ficients of a general model so as to arrive at a simpler model.

5.2 THE PRICE OF FOOD AND THE DEMAND FOR


MANUFACTURED GOODS IN INDIA
Much of the economic literature on India concerns the analysis of the
interplay of agriculture and industry in economic development. Here we
shall investigate one particular contribution to this wider set of issues:
namely, Krishnaji's (1992: 96) argument concerning the perceived narrow-
ness of the market base for manufactured goods in the context of wide-
spread poverty. Krishnaji argued that 'other things remaining the same,
rising cereal prices depress the demand for manufactures' (ibid.: 105). It is
this hypothesis we seek to test in this section with Krishnaji's data (avail-
able on diskette in the data file KRISHNAJI). However, we shall use a
somewhat different specification from that employed by Krishnaji.
In its most general form, our model consists of a constant elasticity
function which features the per capita demand for manufactured consumer
goods (M) as dependent variable as a function of nominal disposable
income per capita (Y), the price index of manufactured goods (P m) and
the price index of cereals (P1). Our model specification thus becomes:

(5.1)
166 Econometrics far developing countries
where E is the error term subject to the usual set of assumptions of the
classical normal linear regression model.
Taking logarithms of both sides yields:
(5.2)
where 13 1 = logA.
Economic theory suggests that 13 2 > O and 134 < O: that is, the income
elasticity is positive while the own price elasticity is negative. As to the
expected sign of 13 3, the coefficient of the price of food, Krishnaji (1992:
106) argues that this coefficient will be negative for the following reason:
if, as seems to be the case, food consumption levels are either inade-
quate for survival (as among the poor) or unsatiated (whether in
quantity or quality, as among sorne above the poverty line), what deter-
mines the allocation process for the majority of the population is not
the total but the 'residual income': that part of income which is left
over after food articles have been bought.
Consequently, if Krishnaji's hypothesis is correct, we expect that 13 3 < O:
the rise in the price of food will adversely affect the demand for manu-
factured goods.
Furthermore, unlike Krishnaji, we assume there is no 'money illusion'
- that is, if nominal income and both price indices rise proportionally, the
demand for manufactured goods remains unchanged. Hence, our assump-
tion - which we shall not formally verify at this juncture - requires that
the slope coefficients in equation (5.2) add up to zero (see Box 5.1). That
is, we assume that equation (5.2) is homogeneous of degree zero, hence:
(5.3)
or
(5.4)

Box 5.1 Money illusion in the demand function


Suppose that income and all prices increase by a constant multiple
k. The demand for manufactures becomes:
logM' = 13 1 + 13 2 log (kY) + ¡33 log (kP1) + 13 4 log (kPm) +E

= log M + (13 2 + 13 3 + 134) log k


The absence of money illusion means that the equiproportionate
rise in income and prices leaves demand unchanged, i.e. M' = M. It
is readily apparent that for this condition to be met it must be the
case that 13 2 + 13 3 + 13 4 = O.
Partial regression 167
Substituting equations (5.4) into equation (5.2), yields:
logM = ~1 + ~z(logY - IogP,,J + ~3(IogP¡ - IogPm) + E

or, alternatively:

log M = /31 + {32 log (;,) + {33 log (~) (5.5)

which conveniently reduces our initial four-variables model to a three-


variable case. In section 5.6 we shall investigate whether the assumption
of no money illusion is valid for these data.

Ignoring the income variable: simple regression


Our main concern is the sign of the coefficient of the relative price of
food vis a vis manufactured goods. Should we then start by ignoring the
income variable and simply regress the demand for manufactured goods
on this relative price variable? If we do this, we obtain the following
results for the period from 1960-1to1980-1 (standard errors in brackets):

log M = 4.59 + 0.266 log ( ~) (5.6)

(0.36) (0.195)
R2 = 0.09, TSS = 0.4158, RSS = 0.3788, ESS = 0.0037
Simple regression (double-logarithmic)
o

o
o

0.260046
Relative price of food

Figure 5.1 Manufacturing demand versus relative food price


168 Econometrics far developing countries
For reasons explained below we shall give detailed information about the
total, the residual and the explained sums of squares of all regressions in
this section. The scatter plot along with the resulting regression is depicted
in Figure 5.1. To say the least, the results are rather disappointing: the
slope coefficient is insignificant and has the wrong sign, and the coeffi-
cient of determination is disappointingly low. It appears, therefore, that
the relative price of food is of little significance in the demand for manu-
factured goods. But the problem with this regression is that we did not
really test Krishnaji's hypothesis at all. His argument was that, other things
being equal, the demand for manufactured consumer goods varies
inversely with the price of food. But, in this case, other things did not
remain constant: in particular, our real income variable, YIPm varied from
year to year. In development research, we do not control the variation of
the different variables and we cannot hold sorne of them constant while
investigating the effect of one explanatory variable on the dependent vari-
able. So how do we deal with this problem?

Removing the linear influence of the income variable: partial regression


To investigate the relation between the demand for manufactured con-
sumer goods and the relative price of food we need to find a method
which allows us to account for the influence exerted by the other explana-
tory variable - the real income variable. How can we do this? The trick
is to remove the linear influence of our real income variable on both vari-
ables under study. In other words, if we cannot control real income, at
least we can try to account for its influence by removing the effect of its
linear covariation with both other variables.
To do this, we run two regressions which both feature the real income
variable as the explanatory variable:

log M = 3.3 + 0.74 log ( ; ) + e12 (5.7)


(0.14) (0.08) m

R2 = 0.82, TSS = 0.4158, RSS = 0.0750, ESS = 0.3408


where e12 refers to the residual obtained by regressing the dependent vari-
able (denoted with subscript 1 since it is the first variable to appear in
model specification (5.5)) with the real income variable (the second vari-
able in the model; hence, subscript 2). The second regression becomes:

log (--Jf) = - 1.06 + 0.57 log ( ; ) + e32 (5.8)


m (0.28) (0.166) m

R2 = 0.38, TSS = 0.5228, RSS = 0.3227, ESS = 0.2001


where e32 are the residuals obtained by regressing the relative price vari-
able (the third variable to appear in model specification (5.5): hence,
Partial regression 169
subscript 3) on the second variable, the real income variable. This last
regression is called an auxiliary regression in as much as it regresses one
explanatory variable on the other.
The residuals e 12 and e32 represent, respectively, the variation in the
demand variable (M), on the one hand, and in the relative price variable
(P/P m), on the other, after removing the linear influence of the real
income variable (Y/Pm) from both these variables. What happens if we
now regress e12 on e32? This is called the partial regression of log(M) on
log(P/Pm) after taking account of the linear influence of log(Y/Pm). A
partial regression, therefore, is a simple regression which features the
residuals of prior regressions as its variables.
The partial regression now gives us the following results:
é 12 = -0.38 e32
(0.067) (5.9)

R 2 = 0.61, TSS = 0.0750, RSS = 0.0288, ESS = 0.0462


This regression does not feature a constant term. Why is this? The reason
is easy to understand: the point of means is always on the simple regres-
sion line, and, in the case of a partial regression, both variables are
residuals which have zero means. Consequently, the partial regression line
goes through the origin which is its point of means.
Interestingly, the slope coefficient in the regression (5.9) now turns out
to be negative and is statistically significantly different from zero. This is
what we would expect, given our initial hypothesis that there exists an
inverse relation between the demand for manufactured goods and the price
of food, other things being equal. But, in fact, we did not hold anything
constant while estimating this inverse relation. Instead, what we did was to
remove the linear influence of the income variable on both the demand
variable and the relative price variable. In this way we managed to bring
out a relation which was hidden deep in the structure of our data.

The partial regression plot


The partial regression (5.9) also provides us with a powerful analytical
graph - a partial regression plot - which enables us to look at what lies
behind a multiple regression coefficient. We do this simply by plotting e12
against e32 , along with the simple regression line between them, as shown
in Figure 5.2. Take a good look at this graph and compare it with the
earlier simple scatter plot in Figure 5.1. The simple scatter plot does not
show any relation between the demand variable and the food price vari-
able. Obviously, the reason is that money incomes also varied during the
period shown. The partial regression plot is more revealing, however, since
it depicts the relation between demand and the price of food after
removing the linear influence of the income variable. Hence, if food prices
170 Econometrics for developing countries
Partial regression (double-logarithmic)

0.132003 o
o

o o
o
o
o o
.E o
"O o o o
e
<1l
E
Q)
o
o o o o
o
-0.98754
-0.240881 0.260917
Relativa price of food
s1a1a'M

Figure 5.2 Partial regression plot of demand on food price

went up, but so did money incomes, the partial regression plot would
remove this covariation between both explanatory variables and only
depict movements in the price variables over and above changes in the
income variable. In the process, nothing is held constant, but neverthe-
less the linear influence of the income variable is removed from the stage
to allow us to look deeper into the patterns of covariation.

Lessons: partial versus multiple regression


Let us now jump ahead a bit and ask the obvious question. How does this
partial regression relate to the multiple regression of the three variables?
In section 5.3 we shall discuss how such a regression can be obtained using
the least squares principle. But at this juncture let us compare our partial
regression with the multiple regression. If we apply the least squares
principle to the model specification (5.5), we get the following results, the
interpretation of which are analogous to those of simple regression:

log M = 2.89 + 0.95 log ( : )- 0.38 log


(0.11) (0.06) m (0.065)
(-if)
m
(5.10)

R 2 = 0.93, TSS = 0.4158, RSS = 0.02877, ESS = 0.3870


Partía/ regression 171
Note that the slope coefficient of the relative price variable in the multiple
regression is exactly the same as the slope coefficient obtained in the
partial regression (5.9). This gives us insight into the way multiple regres-
sion coefficients single out the separate effect of each explanatory variable
on the dependent variable in a context where we cannot hold other things
equal. In regression analysis, therefore, we investigate relations between
two variables by removing the linear influence of third variables deemed
important. In this way we account for the inftuence of the latter without,
however, keeping them constan t. A multiple (or partial) regression coef-
ficient, therefore, cannot be interpreted as the impact of the explanatory
variable in question on the dependent variables, other things being equal,
but instead measures its impact after accounting for (i.e. removing) the
(linear) influence of the third variables.

Decomposing the sums of squares across regressions


Finally, it is instructive to see how the multiple coefficient of determina-
tion, the R 2 in equation (5.10), can be derived from the sequence of simple
and partial regressions. To see this, take another look at equation (5.7)
which depicts the simple regression of log(M) on Log (Y!Pm)· In this case,
the total sum of squares can be decomposed as follows:
TSSS = ESSS + RSSS
0.4158 = 0.3408 + 0.0750
where the subscript s refers to the simple regression. The residuals of this
simple regression then feature as a dependent variable in the partial
regression (given by equation (5.9)) with the following breakdown of its
total sum of squares:
TSSP = ESSP + RSSP
0.0750 = 0.0462 + 0.0288
where the subscript p refers to the partial regression. Note that RSSs =
TSSP, which must be so since the residuals of the simple regression feature
as the dependent variable in the partial regression. Consequently, the
residual sum of squares of the simple regression will be equal to the total
sum of squares of the partial regression.
Now, ESSs is the explained variation in logM due to the inclusion of
the income variable only, and ESSP gives us the added explained varia-
tion due to the further inclusion of the price variable (after removing the
linear influence of the income variable). We obtain the explained varia-
tion due to the influence of both explanatory variables by adding both
these explained sums of squares as follows:
ESSm = ESSs + ESSP
= 0.3408 + 0.0462
= 0.387
172 Econometrics far developing countries
where the subscript m refers to the multiple regression. Indeed, as you
can easily verify, the ESS of the multiple regression 5.10 equals (TSSm -
RSSm) = (0.4158 - 0.02877) = 0.387.
The multiple coefficient of determination can now be calculated as
follows:
R 2m = ESSm / TSS
= 0.387 / 0.4158
= 0.93
which, as can you can see, equals the value of R 2 of the multiple regres-
sion (5.10). In sum, while the simple R 2 gives us the goodness of fit due
to a single explanatory variable, the multiple R 2 gives the goodness of fit
due to two (or more) explanatory variables. The simple and the multiple
R 2 s are directly comparable because the respective regressions feature the
same dependent variable. Hence, in this case, we say that the goodness
of fit increases from 0.82 to 0.93 as we move from the simple to the
multiple regression.
But what is the meaning of the R 2 in the partial regression (5.9)? This
R 2 is not comparable with either the simple or the multiple R 2 s because
it does not feature the same dependent variable. Its dependent variable
is the residuals obtained by removing the linear infiuence of the income
variable on the demand variable. This R 2 is the partial coefficient of deter-
mination which measures the contribution of the relative price variable
in explaining the variation in the demand variable after removing the
linear infiuence of the income variable from both these variables. In other
words, it gives us the added contribution of the price variable to the
multiple regression over and above the infiuence already exerted by the
income variable.
In our example, the simple R 2 in equation (5.6) equals 0.09, which tells
us that, taken on its own, the price variable contributes little towards
explaining the variation in the demand variable. This is not surprising since
the infiuence of a price variable on demand can hardly be meaningfully
assessed without taking account of income variation. In contrast, the par-
tial R 2 equals 0.61 (as obtained in equation (5.9)). This tells us that once
the variations in income are accounted for, the price variable proves to be
important in explaining the remaining variation in the demand variable.

Sweeping out
This section has introduced you to a powerful concept, partial regression,
which allows you to look deep into the structure of data so as to bring
to the surface patterns within the data which are not immediately obvious
but indicative of deeper relations among variables. Partial regression uses
residuals of prior regressions with the raw data. For this reason we used
terms such as 'accounting for the infiuence of other variables' or 'removing
Partial regression 173
the linear influence of other variables'. Or, equivalently, we say that we
'control for other variables' while looking at the relation between any two
variables. These expressions are quite cumbersome to use. Perhaps you
will agree with us that EDA's more colourful expression of sweeping out
makes the point more vividly (Emerson and Hoaglin, 1985). Hence, when
we return to these important concepts of partial regression, partial corre-
lation and partial regression plot in section 5.3, we shall frequently say
that we look at the relation between Y and X, while sweeping out Z,
meaning that we control for the linear influence of Z on both Y and X.

Exercise 5.1
From SOCECON, select the variables Birth (B, the birth rate ), the loga-
rithm of GNP per capita (log Y), and Urban (U, the urban as percentage
of total population). Compare the partial regression of B or log Y,
sweeping out U, with the simple regression of B or log Y. Next, look at
the partial regression of B or U, sweeping out log Y and compare it with
the simple regression of B on U. What do you conclude?

5.3 LEAST SQUARES AND THE SAMPLE MULTIPLE


REGRESSION LINE
As in simple regression, the least squares principie can be applied to derive
an average relation between Y and a number of explanatory variables,
~, j = 1 ... k (on notation, see Box 5.2). To start with, let us consider
the case of three-variables: a dependent variable Y and two regressors,
X 2 and X 3 • The three-variable case is instructive since it provides us with
all the flavour of multiple regression without the excessive strain of
cumbersome mathematical expressions when using summation signs. The
results can subsequently be extended easily to the k-variable case.
In the three-variable case, the sample multiple regression can be written
as follows:
(5.11)
where b1, b 2 , b 3 are the regression coefficients and the e;s are the resid-
uals. Following the least squares principie, the estimates of the regression
coefficients are obtained by minimising the sum of squared residuals:
L e;2 = L (Y; - b1 - b 2X 2 ; - b 3 X 3 ;) 2
with respect to b1, b 2 , b3 •
The first-order conditions yield the following normal equations:
L L X 2 ; + b3 L X 3 ;
Y; = nb 1 + b 2 (5.12)
L Y ;X2; = b1 L X 2; + b2 L A1; + b 3 L X 2;X3; (5.13)
L Y;X3¡ = b1 L X 3; + b2 L X 2; X 3; + b3 L X5; (5.14)
174 Econometrics for developing countries

Box 5.2 A note on notation


In this text, we denote explanatory variables by Xj: j = 2 ... k. We
number the explanatory variables from subscript 2 onwards. One
reason for this notation is that it allows us to refer to the depen-
dent variable as the first variable in the equation. Alternatively, we
can consider Xli = 1, for all i = 1, 2, 3, ... , n, as the variable repre-
senting the constant term.
Hence, when we refer to a three-variable regression, we mean a
regression between Y and two explanatory variables, X 2 and X 3 , with
slope coefficients b 2 and b 3 , respectively. The coefficient b 1 refers to
the intercept.
In sorne applications, however, we also need a subscript for the
dependent variable. In general, when the dependent variable is
denoted by Y, we prefer to use the subscript Y as well. In examples,
however, the dependent variable is often denoted with a different
symbol (such as M in our earlier example), in which case we use
subscript 1 to refer to the dependent variable.

Before deriving the estimators by solving these three equations for b 1,


b 2 and b 3 , let us first take a closer look at each equation in turn because
each reveals sorne important properties of the estimated regression.

The mathematical properties of the least squares regression line


Equation (5.12) can be rewritten in two different ways, each of which
highlights a different property of the least squares line. First, dividing both
sides by the number of observations, n, yields the following result:
- - -
Y¡= b1 + b2 X 2 + b3 X 3 (5.12a)
which tells us that the point of means satisfies the least squares regres-
sion equation. In other words, the least squares regression line passes
through the point of means, a result which is familiar from simple regres-
sion.
Second, rewriting the normal equation (5.12) as follows:
L (Y¡ - b1 - b 2 X 2; - b 3 X 3;) =Le;= O (5.12b)
gives us the familiar zero-mean property of the least squares residuals.
Third, the normal equations (5.13) and (5.14) can be rearranged as
follows:
(5.13a)
and
Partial regression 175

(5.14a)
which show that the least squares residuals are uncorrelated with each of
the regressors, a result which is again familiar from simple regression
analysis.
Fourth, the least squares regression is a plane in the three-dimensional
space, though it is often referred to as the 'regression line' by analogy
with the case of simple regression. The resulting regression line which
yields the average relation of Y for given X 2 and X 3 is written as:
(5.15)
Now, equations (5.13a) and (5.14a) imply that the predicted (or fitted) Y
values are uncorrelated with the residuals. This can be shown as follows:
(5.16)
Hence, the mathematical properties of the least squares regression line
in multiple regression are a simple extension of the properties of the
simple regression line.

Multiple regression coefficients


The solutions of the simultaneous normal equations for the sample regres-
sion coefficients yield the least squares estimators for b1, b 2 and b 3 • Since
the resulting formulae become somewhat cumbersome, it is useful to
simplify the algebraic expressions by using the following notational
conventions:
SY2 = L Y¡X2 ; - n YX2
Sy3 =L Y¡X3¡ - n YX3

First, the intercept, b 1, can be obtained from equation (5.12a) which


specifies that the regression line goes through the point of means, as
follows:
- - -
b 1 =Y¡ - b 2 X 2¡ + b 3 X 3¡ (5.17)
The coefficient b 2 is given by,
b = S33Sy2 - S3ySz3
2 (5.18)
S22S33 - (S23)2

It is important to note that b2 depends not only on X 2 but also on the


other explanatory variable X 3 since the formula in equation (5.19) also
includes the terms S33 and S23 which depend on the sample values of X 3 •
176 Econometrics far developing countries
It is instructive to divide both the numerator and the denominator in
(5.18) by the product term (S22 .S33) in arder to obtain the following alter-
native expressions for b2 :

b = bY2 - bY3b32 = rY2 - rY3r23 Sy


(5.18a)
2 1- b23b32 1- ?z3 S2

where, byj = the slope coefficient of the simple regressions of Y on Xj,


for j = 2,3; b;j = the slope coefficients of the simple regression of X; on
Xj, i,j = 2,3, i i= j; r;j = the simple correlation coefficients among Y, X 2
and X 3 , for i,j = Y,2,3; s; = the sample standard deviations of Y, X 2 and
X 3 , for i = Y,2,3.
Similarly, the corresponding formulae for the coefficient b3 are as
follows:

(5.18)

b = bY3 - bY2b23 = rY3 - T12T23 Sy


(5.19a)
3 1- b23b32 1- ?z3 S3

Expressions (5.18) and (5.18a) reveal that b2 depends not only on the
slope coefficient of the simple regression of Y on X 2 , but also on the
slopes of the simple regressions of, respectively, Y on X 3 and X 2 on X 3
(the auxiliary regression). A similar argument can be made for b 3• Hence,
in general, simple and multiple regression do not yield the same estimates
of slope coefficients of simple and multiple regressions. There are,
however, two exceptions to this general rule, about both of which it is
instructive to know. To see what these exceptions are, we suggest you
attempt exercise 5.2 befare reading on.

Exercise 5.2
Using equations (5.18) or (5.18a) and (5.19) or (5.19a), show that the slope
coefficient of X 2 will be the same in the simple regression of Y on X 2 and
in the multiple regression of Y on X 2 and X 3 , if (a) X 2 and X 3 are uncor-
related with each other; and (b) b3 , the multiple regression coefficient of
X 3 , equals O. Generate a data set to illustrate both of these special cases.

The proofs are simple and straightforward. Each case, however, gives
us sorne interesting insights into the question as to how multiple regres-
sion seeks to disentangle the separate effects of different explanatory
variables on the dependent variable. We discuss each case in turn.
Partía! regression 177
Orthogonality and perfect collinearity of regressors
If X 2 and X 3 are uncorrelated, b 23 , b 32 and r23 all equal O, and, hence, the
simple regression coefficient of Y on X 2 equals the slope coefficient of X 2
in the multiple regression of Y on X 2 and X 3: that is, b2 = by2 • In this case
we say that the regressors X 2 and X 3 are orthogonal.
Intuitively, this result makes sense. If the explanatory variables in a
multiple regression do not covary linearly with one another, it follows that
multiple regression analysis allows us to distinguish clearly between their
separate effects on the dependent variable. Multiple regression, therefore,
becomes the simple addition of the constituent simple regressions. Sweeping
out is unnecessary because the explanatory variables do not overlap.
The opposite of orthogonality occurs when a perfect linear relation
exists between X 2 and X 3 , that is, X 2 = a + b.X3 , where a and b are non-
zero constants. In this case, r23 = 1: we say that both regressors are
perfectly collinear. From equation (5.19) it follows that b 2 will be inde-
terminate because the denominator in the formula will be zero: we can
either regress Y on X 2 or Y on X 3 but not Y on both X 2 and X 3 together.
In many regression packages you may get a message such as 'singular' or
'near singular matrix' when attempting to perform a regression. This
message means that sorne of your regressors have a strong linear rela-
tionship with one another and so, since it involves a division by zero, the
computer cannot complete the calculation.

Adding a superftuous variable


If b 3 = O, dropping X 3 from the multiple regression will not affect the
slope coefficient of X 2 • That is, b 2 = by2 • The proof is straightforward: if
b 3 =O, and X 2 and X 3 are not perfectly collinear, it follows that the numer-
ator in equation (5.19) (or (5.19a)) equals zero, which then yields an
expression which can be substituted into equation (5.18) (or (5.18a)) to
show that b 2 = byz. Intuitively, this makes sense as well. If the addition
of X 3 does not add anything in terms of explaining the variation in Y,
dropping X 3 from the regression model should not affect our results. In
this case, X 3 is a superfluous variable: it contributes nothing once we have
already taken account of the inftuence of X 2 •
Note, however, that b 3 = O does not mean that bY3 = O as well. Taken
on its own, X 3 may covary with Y and, hence, the slope coefficient, bY3,
in the simple regression is not necessarily equal to zero. Once X 2 enters
the scene, however, X 3 is effectively eclipsed by the introduction of this
new variable and, hence, its influence dwindles to zero. This teaches us
an important lesson. Whether, in a multiple regression, an explanatory
variable turns out to be superfluous or not cannot be deduced from its
performance in its simple regression with the dependent variable; a vari-
able which plays no role at all in the true model may appear significant
if we begin from a more specific, incorrectly specified model.
178 Econometrics for developing countries
Extension to the k-variable case
The analysis above can easily be extended to the k-variable case. The nor-
mal equations obtained by minimising the sum of squared residuals have a
similar structure to the one shown in the case of three variables. In this
case, however, the solution for the bj (j = 1, 2, 3, ... k) in terms of the Y;
and Xji values becomes cumbersome when using summation signs once we
consider the case of k > 3. The use of matrix algebra is more appropriate
to derive the general solution for this system of equations. In practice, of
course, we rely on computer software to salve the system of equations
involved. Our interest here, however, is with the general properties of the
least squares solution. Box 5.3 provides a summary of these properties.

Box 5.3 The properties of the least square line

1 The least squares line passes through the point of means.


2 The least squares residuals have zero mean.
3 The least squares residuals are uncorrelated with and, hence,
orthogonal to each of regressors.
4 The least squares residuals are uncorrelated with and, conse-
quently, orthogonal to the fitted values of the dependent variable.

The concepts of orthogonality and perfect collinearity can equally be


extended to the k-variable case. The case of orthogonality is straightfor-
ward: Xj is said to be orthogonal to all other explanatory variables if it
is uncorrelated with each of the other regressors included in the model.
In this case, the simple regression coefficient of Y on Xj will be the same
as the slope coefficient of Xj in the multiple regression of Y on Xj with
all other regressors included in the model.
As far as the question of collinearity is concerned, its generalisation
to the k-variable case is somewhat more complex. If a set of constants
aj (j = 1, 2, 3, ... k), not all equal to zero, can be found such that

(5.20)
then X 1 (i.e. the constant term: X 1; = 1, for all i), X 2 , X 3 , ••• , Xk are said
to be perfectly multicollinear because, in this case, any of the Xjs can be
expressed as a linear function of the remaining regressors. As in the three-
variable case, it will then not be possible to salve the system of normal
equations and, hence, multiple regression breaks down.
The question of superfluous variables extends equally to the k-variable
case in multiple regression. If we find that bj = O, dropping Xj from the
regression model will not affect the other coefficients in the model. But
this does not mean that this variable ~ will yield zero coefficients in
Partial regression 179
regressions (simple and multiple) which involve subsets of the initial
broader model. In other words, if regressing Y on X, Z and W yields a
zero coefficient for X, it does not necessarily imply that the slope coeffi-
cients for X will also be zero in the simple regression of Y on X, in the
regression of Y on X and Z, or in the regression of Y on X and W.
Orthogonality or perfect collinearity as well as slope coefficients which
yield exact zero values are in fact extreme situations. With respect to
collinearity, as far as regression is concerned, life is simplest when regres-
sors are orthogonal and it is impossible when perfect collinearity prevails.
Most socioeconomic data, however, fall somewhere in between these two
extremes: regressors derived from non-experimental data tend to covary
with one another but not in an exact linear fashion. However, more often
than not the strength of the correlation between regressors tends to be very
high, which makes it difficult to distinguish clearly between their separate
effects on the dependent variable. Moreover, due to the presence of
collinearity, a variable may perform well in either simple or multiple regres-
sions until sorne other variable takes the stage and renders it superfluous.
Interpreting multiple regressions, therefore, is a complex problem since
the presence of imperfectly collinear regressors blurs the picture when
dealing with non-experimental data. The main lesson we learn from
looking at the algebra of least squares is that the inclusion or deletion of
one or more explanatory variable(s) from a regression will generally not
leave the regression coefficients of the other regressors included in the
model unchanged. In other words, when interpreting the slope coefficient
of a particular regressor, Xj, it usually matters which other regressors are
also included in the model. That is, the estimated coefficient of any
regressor can change quite dramatically across a range of specifications
which differ depending on which other variables are included in or
excluded from the model. We shall return to this issue in section 5.7.

The coefficient of determination


The coefficient of determination, the square of the multiple correlation
coefficient, measures the strength or goodness of fit of the combined influ-
ence of all regressors on the dependent variable. It is defined as the ratio
ESS!RSS - the proportion of total variation in Y explained by the set of
the explanatory variables Xj (j = 2, 3, ... , k). In multiple regression, these
sums of squares are obtained as follows:
TSS = Syy = ESS + RSS
where
ESS = explained sum of squares = ~ bjSYj
1
RSS = residual sum of squares = ~ e] (5.21)
i

where Syj = ~ Y;X¡; - nYXj; j = 2, 3, ... , k


¡
180 Econometrics far developing countries
As with simple linear regression, therefore, the total sums of squares, TSS,
can be divided up into two components: the explained sum of squares,
ESS, which gives the variation in Y explained by the variation in all regres-
sors included in the model, and the residual sum of squares, RSS. The
coefficient of determination is then obtained as follows,
R 2 = ESSITSS (5.22)
which measures the goodness of fit for the multiple regression as a whole.
The R 2 can also be obtained by computing the square of the simple coef-
ficient of correlation between Y; and its predicted values derived from the
multiple regression. As explained above, the R 2 of a multiple regression
measures the explained variation in Y due to all explanatory variables
included in the model as a proportion of the total variation in Y. In
general, however, it is not possible to disentangle the ESS with respect
to the contribution of each of the regressors. In other words, the ESS of
a multiple regression is not a simple addition of the R 2s of the set of
simple regressions of the dependent variable with each of the regressors
in the model. The reason is that, in general, regressors covary with one
another and, hence, they overlap in terms of explaining the variation in
Y. Our example in section 5.2 showed this point very clearly.
If, however, the regressors in a multiple regression are orthogonal, the
R 2 of the multiple regression is the simple addition of the R 2s of the corre-
sponding simple regressions. In practice, exact orthogonality seldom
occurs, but if regressors are near-orthogonal, the sum of the simple R2s
will not be much larger than the multiple R 2• If, in contrast, a variable Xj
turns out to be totally superfluous in a multiple regression and hence
bj = O, dropping it from the model will not affect the ESS and so the R 2
will also be unchanged. We seldom get bj exactly equal to zero, but a
regressor can be said to be redundant for practica! purposes if dropping
it from the regression hardly affects the ESS.

5.4 PARTIAL REGRESSION AND PARTIAL CORRELATION


We now take a closer look at partial regression - the technique of
sweeping out the influence of other regressors when analysing the rela-
tion between the dependent variable and one specific explanatory variable.
In the three-variable case, the partial regression of Y on X 2 is the simple
regression of the residuals of, respectively, Y on X 3 and X 2 on X 3•

Partial and multiple regression coefficients


To avoid an overdose of cumbersome notation, let y, x 2 and x 3 denote the
deviations of, respectively, Y, X 2 and X 3 from their respective means.
Consequently, the slope of the partial regression of Y on X 2 which we
denote here as b/ is given as follows (ignoring subscripts i),
Partial regression 181

:¿(y - b13X3)(x2 - b23X3) :¿y(x2 - b23X3)


bl = (5.23)

since x 3 is uncorrelated with the residuals (x 2 - b23x 3). Substituting the


formula for b23 into equation (5.23) yields:

2:YX2 2:x~ - 2:YX3 2:xzX3


bl = = b2
2:xi 2:x~ - C2:xzX3)
2

which is identical to the slope coefficient of X 2 in the multiple regression


of Y on X 2 and X 3 as given by equation (5.18).
Similarly, b3 can be shown to be equal to the slope of the partial regres-
sion of Y on X 3 , after sweeping out X 2 • Hence, partial regressions yield
exactly the same slope coefficients as the corresponding coefficients in the
multiple regression. This explains why the regression coefficients in the
multiple regression are most commonly referred to as partial regression
coefficients.

The coefficient of partial correlation


The coefficient of correlation between (y - by 3 x 3) and (x 2 - b23 x3), the
residuals from the regressions of Y on X 3 and of X 2 on X 3 , is called the
coefficient of partial correlation. We shall denote this partial correlation
as ryz 3 : the subscripts before the dot '.' indicate the two variables under
consideration and the number after the '.' indica tes the other regressor
the linear infiuence of which has been removed. The coefficient of partial
correlation, therefore, can be expressed as follows:

(5.24)

which, after sorne manipulation, can be simplified as follows:

(5.25)

In other words, the partial correlation between Y and X 2 is nothing but


the simple correlation between both variables net of the correlations
between Y and X 3 , and X 2 and X 3 • In a similar fashion, the partial corre-
lation between Y and X 3 is obtained as follows:
182 Econometrics for developing countries

(5.26)

Partial regression and correlation in the k-variable case


The concept of partial regression can easily be extended to the general
case of k variables. The partial regression between the dependent vari-
able and any of the explanatory variables involves sweeping out the
combined infiuence of all other regressors in the model, which is done by
regressing Y and the explanatory variable in question on all other regres-
sors so as to obtain two sets of residuals which then feature as the variables
in the partial regression. The auxiliary regressions are thus multiple regres-
sions (see Box 5.4). As in the three-variables case, the slope coefficient
of this partial regression will be equal to the corresponding slope coeffi-
cient in the multiple regression.
As far as the concept of partial correlation is concerned, it is useful to
consider the notion of the 'arder of partial correlation'. A partial corre-
lation which involves sweeping out only one regressor (as was the case of

Box 5.4 The concept of auxiliary regressions

1 Auxiliary regressions involve regressions between two or more


explanatory variables.
2 In the multiple regression of Y on X 2 and X 3 , the auxiliary regres-
sions are the simple regressions of X 2 on X 3 and of X 3 on X 2 •
These regressions are useful for two purposes: first, in partial
regression analysis they allow us to sweep out the linear infiu-
ence of each regressor on the other, and, second, they show the
extent to which both regressors are collinear or not.
3 In the k-variable case (k > 3), the auxiliary regressions are
multiple regressions since here we regress each explanatory vari-
able on all the other regressors included in the model. Again
these regressions are useful for two purposes. First, they allow
us to sweep out the linear infiuence of all other regressors on
the remaining one. Second, they reveal the extent of multi-
collinearity among the regressors. The reason is that pairwise
simple regressions and correlations between regressors are not
sufficient to check for the presence of multicollinearity. There is
indeed the possibility that three or more variables covary in a
pattern which is not revealed by simple correlations between any
pair of these regressors.
Partial regression 183
ry2.3 above) is called a first-order partial correlation. Similarly, a partial
correlation obtained after sweeping out two regressors is called a second-
order partial correlation, and so on. The simple correlation between any
two variables - far example, r y¡, j = 2, 3, ... , k, can be then looked upan
as a special case of a zero-order partial correlation. Obviously, in the
k-variable case (which involves a dependent variable and k - 1 explana-
tory variables) we are most commonly interested in the partial correlations
of arder k - 2 which give us the individual contributions of each regressor
in terms of reducing the residual sum of squares obtained by regression
Y on all other regressors.

The partial regression plot


The partial regression plot - also called the added-variable plot - is a
simple yet powerful graphical tool of analysis which allows us to look
carefully at the scatter plot that underscores a particular regression coef-
ficient in a multiple regression. As we saw above, a partial regression plot
is a scatter plot featuring residuals on its axes rather than the original
variables. In the k-variable case, there are k - 1 partial regression plots
each of which throws a different light on the structure of the data.
It is customary to plot the residuals along with the partial regression
line which goes through the origin, since residuals have zero mean. The
plot, therefare, shows how well a slope coefficient of a multiple regres-
sion is supported by the data. Partial regression plots can be interpreted
in much the same way as you would look at a simple scatter plot (see
Chapter 3). They allow you to look far non-linearity, heteroscedasticity,
outliers and infiuential points. There is, however, an important difference:
partial regression plots depict residuals and, hence, they allow you to look
deep into the structure of the data by controlling far the combined infiu-
ence exerted by other regressors. Consequently, a partial regression plot
depends on the other regressors taken into consideration and so the plot
will differ depending on which other regressors are included in the model
and on any transfarmations applied to the variables in it.

Exercise 5.3
In section 5.2, we studied the effect of the price of faod on the demand
far manufacturing goods in India in the context of a three-variable model
arrived by imposing the assumption of 'no money illusion' on the general
model given by equation (5.2). In this exercise, you are requested to drop
this assumption and to use model specification (5.2) instead to work out
the fallowing questions:
1 Regress log(M) on log(Y), log(P1) and log(Pm).
2 Compute the three partial regressions and their coefficients of partial
correlation.
184 Econometrics far developing countries
3 Compare the partial regression plots with the corresponding simple
scatter plots.
4 How would you interpret, respectively, the simple, partial and multiple
coefficients of determination?
5 Show how the multiple R 2 can be derived from the ESSs of a hier-
archy of simple regressions involving the raw data as well as residuals
from prior regressions.
6 Compare the results of the four-variable model with those of the three-
variable model based on the assumption of 'no money illusion'.
7 In your opinion, do you think the assumption of 'no money illusion'
is warranted? Why?

5.5 THE LINEAR REGRESSION MODEL


Up to this point we have dealt with the sample regression line which in
the k-variable case can be formulated as follows:
(5.27)
where the b¡ (j = 1, ... k) are the least squares estimators of the coeffi-
cients !3¡ (j = 1, ... k) of the population regression model which is:
(5.27)
where the ES denote the error terms of the population regression line as
distinct from the es which are the residuals obtained from the estimated
sample regression.

Model assumptions
The classical model is subject to the following usual assumptions:
1 The population regression is adequately represented by a linear func-
tion of the k variables included in the model: E(Y;) = !31 + !32 X 2; +
!33 X3; · · · + !3¡ X¡;;
2 the error terms have zero mean: E( E;) = O;
3 constant variances: V( E;) = a 2 , for all i;
4 and zero covariances: E; and E¡ are uncorrelated for i =/= j;
5 the error term and the X¡s ha ve zero covariances: E( E;,X¡;) = O, for all
i and j;
6 there is no perfect collinearity between regressors: no exact linear rela-
tion exists between the X¡;s, j = 1, 2, 3, ... k, where Xli = 1 for all i
(i.e. the constant term).
Apart from the last one, all assumptions are exactly the same as with
simple regression. An added assumption is needed to ensure that least
squares will yield a solution for the coefficients of the model. The das-
Partial regression 185
sical normal multiple regression model requires that we add the familiar
normality assumption:
7 the error terms have identical normal distributions: E; - N(O,a2),
i = 1 ... n.

Statistical properties
Given the assumptions of classical multiple linear regression, the Gauss-
Markov theorem can be extended to the k-variable case to prove that, as
in the simple regression model, the least squares estimators of 13¡ are
BLUE - best, linear and unbiased estimators. Furthermore, if the
normality assumption is valid, the least squares estimators will also be ML
(maximum likelihood) estimators and, hence, have the property that they
have mínimum variance among all estimators. As with the sample mean
and the least squares simple regression line, the least squares estimators
are unbeatable if the assumptions of classical normal multiple regression
are satisfied in practice. As befare, the normality assumption lays the basis
for statistical inference in the linear regression model.

The variances of the sampling distribution of the b/>


Given these assumptions, the least squares estimators, b¡ (j = 1, ... k),
are also normally distributed since they are linear functions of the Y;s.
Furthermore, E(b¡) = 13¡, since the least squares estimators are unbiased.
The standard errors of the b¡s are given by:
a2 k k --
V(/31) = - + ~ ~ X;X¡ Cov (/3;, /3¡) (5.29)
n i=2i=2

where

for the constant term, and, similarly, for the slope coefficients, we get:
a2
V(/3¡) = ( ) ;j = 2,3, ... ,k (5.30)
S¡¡ 1- R¡2

where, R 2i = the coefficient of determination of the auxiliary regression


of X¡ on all other regressors included in the model.
Equation (5.30) shows that the variance of the slope coefficient of any
of the regressors, X¡, depends on three factors: (a) the variance of the
error term; (b) the sampling variance of the regressor of X¡; and (c) the
coefficient of determination (R 2i) of the auxiliary regression of regressor
j on all other regressors included in the model. Consequently, other things
186 Econometrics far developing countries
being equal, the variance of the regression coefficients will be smallest
in the absence of multicollinearity. To see this, we can rewrite (5.30) as
follows:
a2 1
V(/3) = ~ (l _ R-2) ; j = 2,3, ... , k (5.31)
]] 1

or
a2 .
V(f3j) = - · V!Fj ; ¡ = 2,3, ... , k (5.32)
sjj

where
1
V!Fj =l _ R.2 ;j = 1,2,3, ... ,kl]! (5.33)
1

are called the variance infiation factors of the coefficients of the regres-
sion model. The V!Fjs measure the degree of multicollinearity among
regressors with reference to the ideal situation where all explanatory vari-
ables are uncorrelated (R2i = O implies V!Fj = 1, for all j = 2, ... k). If,
however, R 2i is positive but smaller than 1, V!Fj > 1, approaching infinity
when R 2i approaches l. If we obtain high values of the R 2is, however, we
should not conclude that our regression coefficients will also have high
variances. Much depends on other factors, namely, the sampling variances
of the regressors and the error variance. Por example, a small error vari-
ance coupled with a relatively high degree of multicollinearity can still
produce acceptable sampling variances of our estimates.
Conversely, the prevalence of large estimated standard errors for the
coefficients of the regression model cannot always be ascribed to the
problem of multicollinearity. It is indeed equally possible that our model
was badly specified and, hence, <loes not capture the real determinants of
the dependent variable in question. Consequently, our error variance will
tend to be large. Alternatively, our X variables may vary too little to get
meaningful precision in estimation.

The estimator of the error variance


In practice, we do not know the value of the variance of the error term,
but, as in the case of simple regression, we use the ratio of the sum of
squared residuals of the estimated model to the number of degrees of
freedom (i.e. sample size minus the number of estimated coefficients) as
an estimator for the error variance. Hence:
s2 = RSS!(n - k) (5.34)
where, n = the sample size, and k = the number of coefficients in the
regression line. This expression gives us an unbiased estimator for the
Partial regression 187
variance of the error term, the square root of which is the standard error
of the regression. The value is substituted into equations (5.29) and (5.30)
to obtain estimators far the variances (and standard errors) of the least
squares regression coefficients, allowing statistical inference in multiple
regression. Befare turning to the latter issue, it is important to stress that
the inferences we make from, and the properties we ascribe to, the least
squares estimators are only valid if the assumptions of the classical normal
linear multiple regression model are reasonably valid in practice.

Checking the assumptions of the regression model


As with simple regression, it is useful to make use of analytical graphs
along with numerical summaries as diagnostic tools to check whether the
model assumptions are reasonably satisfied in practice. To check model
assumptions, it is a good strategy to start from the ground upwards; that
is, first take a good look at the empirical distributions of each of the vari-
ables in play. Next, take a look at the two-by-two scatter plots of all
variables. This can best be done with a scatter matrix which arranges all
possible two-way scatter plots between the variables in matrix farm. You
have already come across an example how to do this in the introduction
to this book where we looked at the relation between the crude birth rate
as a dependent variable and GNP per capita and infant mortality as
explanatory variables. In this example, we first looked at the individual
empirical distributions of each of these variables and at their two-way
scatter plots befare arriving at our preferred multiple regression. In the
exercise in the introduction you were asked to redo this example with
data from the file SOCECON. Suppose that we now bring two more
explanatory variables in play: say, the human development index and the
degree of urbanisation as measured by the percentage of the population
living in urban areas. A quick graphical check (not done here) of their
empirical distributions tells you that the latter variable is reasonably
symmetrical in nature but, interestingly, the human development index is
skewed to the left. Figure 5.3 shows the corresponding scatter plot matrix
with the raw data on all the variables in the analysis (GNP per capita,
infant mortality, human development index, the degree of urbanisation,
and the crude birth rate).
The scatter plot matrix is a powerful graphical tool of analysis
which allows you, at a glance, to detect the presence of non-linearities,
heteroscedasticity and outliers or infiuential points across the range of
two-way scatter plots. These two-way plots do not reveal the patterns
inherent in multiple regression since this involves sweeping out the infiu-
ence of third variables on the interplay between the dependent variable
and a particular explanatory variable; but the simple scatter plots are
nevertheless useful since they often warn us against problems which may
crop up again in multivariate analysis. In this enlarged example we faund
4 142 0.191 0.932
o
i & 8 33590

o'O- ~o
o
0

lºi 0
o
i?o<:tJo
80

ll
GNPc
ºº o Sao cPO~
~ ~o
~~Qaaa~
o
O ro
ooJliilr ~CD ób
O o & cPOO
oerner2a ~ ~ ·~º
o o 8
100
o o o
~ 'b o o~ oº ºº
o .Jbeflº
(l)
142 º a i
º~~ ~~~'8
o o n-,~ ºº o o o o o~8°
º~~º
INFANT CP-'&cP o o
MORTALITY &o~~
~
º~~~L
o o o ºº
4 1 &wro~~ 00
o cP 100

lí:;;.,~/~Í>~<»oo°;,
1

~ºq,~ ci>~ <oºº o


éP 8 o
º0::> 00
o ºo o
o Ó' o <óo 0 o o o
~
li~~ i,,:
::So oCb
ººoººo ºªººo
~º8 ~ ~4;º
"Q)@O o
o o o
'Q) º8 ooº o % URBAN o llR-
~oe~
o 0 o
o o
~ o Q>o
, fj
o o
.,¡, oo ! oó>
o ~o
o
Q;i <Q,o o~ §'ºoººº l 5
o o o o

JJ cP<t,~~l'fp
~o

ft''"'~ ~o°ª5b~
0.932 o
o
~~go
ºº ~Ji:. ºª~ ~óC o
ó> o o o o
oº oJ~~~&
o o
HDI ~ ooo
~"" ~
o ºct1ocn
o o 00 u-(JJ o o
~o
0.191 o
º~~~~ºº"~~
o o
oo oºo ol!I
o O"
ºo
<l>
o
54
'6~o

~
o o o o
00
ge §r;{? 'iblt @ o~ º?; o o
~8
o q, o Cb o
g@ó~~º@~~io
0 0
o'a.i/lf 0 '6 o'2i ºº&%>o o 0

:li~aºoºº
o o On0,.
~o'gog;
0~ 0 8°0 BIRTH
o o c8o 6b oo o oº -0'%º o
&,_ 000 o ~ ,!' o 01; :Pai;o
'fl>~
J;:Jº 9>oºó>gg'9@0o
&
oo E- ºo o "U~º& c»&o~m o o o0 cgw
9
100 33590 5 100 9 54

Figure 53 Scatter plot matrix wíth untransformed data


2 11.9164 0.036481 0.868624

~ 'll ~
sº o588.o~ 0 10.422
• ºg o c9 ~o
oºº 0 o00
~~~~o
o o o o
~&o º
0 0
~O~~ ~ 80~~ co
•F~"'º'b
o ºo o
LOG (GNPc) ºo o
~"9~~ o~O-.~J3
~ ºº~o
<D
oo
o o o (l)O? B~!ll>o
4~~~cP o ºº °o°o~Voo o
o ~li)- o o
o o 4.60517
11.9164 o o

º~~o 8 8 ,.,_l_ 'Oº o


o
8~o o
8 ~ ºo oo
oºº ºo
~~~~º
o j 8 o
SR INFANT '069'. o ó>
º~~<()
ºº~o
\
~gé9Joo00 MORTAUTY o -~ºo~ ~ g:io o
0
o§l~
cw~~~f9o
O
9 o ººº oOo

2
O c@~O~~ o ig : o~ •ºo o

c?o 8 ~ ~ ~ºl> ~~o 9:,


100
o o o <o ~o
oº ' o9:, o
o
o tTJgº c9º~ ~q..º§º cf-0 qsf' 0 ~ ºo Rcb ~ oo
D W,ºrJJ O (!r ~ ~~ i
ruP~ u.g@ ° ~) o~~c8 CSº<S (5'10 ó>o

~
i(o®JP& 0°08~ 68~<&
0
o c9 % URBAN 00

ºot~º'I,,º ~ g o~ o~gxp~
o
o oaocoº 00
~ Q %>º~ o
Q)O O 09> 00 -o o o ~o eo-v
o o o 8ºººº1
00 o o 5

~~\ºo
0.868624 fJr&'rJld
tcn:¿s ~ o
o0
rf,.o':o 9'go o
~ ~o

l
0 o o
& o0 oa:o

'~¡i;o 0 ~ <$'
Oo 0 0 0

ó' "b
o o
o
o c9 o o oo
o ºB<b
o <oº
SQUARE HDI o J\q ~
0
0
o
8
~r ºct10<(9~
0 0
o¡/loo oº o oO ºo

0.036481 0oci:i~ -~ º·~ oo º~º ~á;o


o
~
54
%
~
Jiº ºº
o~<el§C-0
oº'
0
§~Jo oo o oo <b~,: o oo 0 º2-. ºº o o
8> ~cP 'O o8 ººcP%> ~u- o~ o
º~8~~~
o 8re1<000
o o o
8 8~~~o 0 0 11I:a~~o ~~O~ 8°0 o -t!lbo~~ o
BIRTH
0 0
C\m
cBO 61 Oo
&o~ o~ 0 c~
~ ºoººo o ~ rÍ8wº o~~~
c8Q> ~~ W'Cb<()
oog
o ºo o <e 6> ~º~@ oo o0 o 0
9
4.605i7 100.422 5 100 9 54

Figure 5.4 Scatter plot matrix with transformed data


190 Econometrics far developing countries
it useful to try out the following transformations: the logarithm of GNP
per capita, the square root of infant mortality, and the square of the human
development index, leaving the other variables (the birth rate and the
degree of urbanisation) unchanged. This yields the scatter plot matrix
depicted in Figure 5.4. which no longer shows majar non-linearities in the
two-ways plots. This scatter plot matrix shows that the problem of non-
linearities has been largely removed from the two-way plots. This is usually
a good indication that it is also better to work with these transformed
variables in multiple regression.
To look specifically at the performance of each regressor in a multiple
regression we use partial regression plots. These are powerful graphical
tools of analysis which help us to assess how well the data points support
the slope of a multiple regression coefficient. To do this you look at each
partial regression plot in turn in much the same way as you would look
at simple scatter plots. These partial regression plots help you to detect
non-linearities, heteroscedasticity and the presence of outliers, leverage
and influential points. Figure 5.5, for example, shows the matrix of partial
regression plots of the regression of the crude birth rate on the logarithm
of GNP per capita, the square root of infant mortality, the square of the
human development index, and the degree of urbanisation. lt is clear from
these plots that the logarithm of GNP per capita and the degree of urban-
isation appear to be of little relevance. While infant mortality and the
development index perform better, the partial regression plots revea} that
the results are undoubtedly affected by the presence of influential points
and outliers.
Residual plots can also be used in much the same way as we used them
in simple regression analysis. Hence, we can plot residuals against the
predicted values of Y. Alternatively, we may also choose to plot the resid-
uals against each of the regressors. lf the residuals are heteroscedastic,
the latter type of plots can be useful to check whether the heteroscedas-
ticity of the residuals is linked with any one of the regressors. For example,
in a regression of household expenditures on manufactured consumer
goods as the dependent variable and, respectively, household income and
household size as independent variables, we often find that residuals are
heteroscedastic. This heteroscedasticity, however, is usually related to the
income variable, and not to the demographic variable. Rich households
not only consume more manufactured goods but also have more varia-
tion in their expenditures patterns. A plot of residuals against the income
variable will revea} how the level and spread of such expenditures both
increase with higher incomes.
Studentised residuals, hat statistics and DFBETAs can easily be
extended to multiple regression along with careful scrutiny of partial
regression plots. The formula for the hat statistics becomes more complex
since we are now dealing with severa} explanatory variables at once (see,
for example, Hamilton, 1990: 130). However, the basic principie for its
{a) coef = 1.1488211. se= 0.8899388. l= 1.29 (b} coef = 2.6955902. se= 0.59670131. t = 4.52
19.8715 o o o
19.87151
o
o
o o

x o o
o
o o <-O ºo
ºo o

o oº
~ ~ºOt:io o x
! o

o 8 Cb
~e
o 'iJf'tPº <JAº% o
'-'Oº
9 .9
-
.et:::: ºº~o o oº o oB>
o o 00
o óh.
~-u
oólo
8
o
ct:Po
fP
o .et::::
:o 0 :o
Q) o o o Q) 1 o
o o o
o o o
o

-26.0901 -,-
-1.46011
o

e( lgnpc 1X)
.71423
--27.9042 L
-2.03139
o

e( sinfmort 1X)
2.81092

(e) coef = -i 9.321726. se= 7.9017212. t = -2.45 (d} coef = 0.0090213. se= 0.04005153 t= 0.23
19.8638 o
o 20.7867 ~o o

o o
o
o
o ~ o o o
ºo o €l R o o o
x x o ºa o -º~~o~ ';.RO:P0 °o o
.et::: o~ o S~" e
~ ~
o 0
.e o o o o Q) & o
t:::: o o o oº & o'() o
:o o

i5 o o 0
ºº &
0
o
Q) o Q)
o o o o o o o 8
o ºo o

-24.3383 o -25.6923 . - - o
-0.18374 0.163167 -37 .8281 34.1522
e( hdi2 IX) e( urbanpop 1 X }

Figure 5.5 Partial regression plots of regression of birth: on log(GNPc) (b) on Infmort05 ; on HDF; and (d) on urban
192 Econometrics for developing countries
derivation remains the same. Since the least squares estimators are linear
functions of the Y¡s, it follows that the predicted Y;s are also linear func-
tions of the Y;s. The hat statistic hu is the coefficient of Y; in the function
which expresses the predicted Y; as a linear combination of the observed
Y;s. It can be shown that this statistic measures the distance a particular
configuration of X values corresponding to data point i lies in relation to
the main body of the X data. Hence, hu measures leverage as a result of
the position of various Xjs in combination. The studentised residuals as
well as the DFBETAs are calculated analogously to the case of simple
regression. Note, however, that in multiple regression we have a set of k
DFBETAs, one for each coefficient in the regression. In a three-variable
case, for example, a particular data point may exert infiuence on the slope
coefficient of X 2 without affecting the slope of X 3•
Obviously, a critica! assumption of the classical linear regression model
is that all relevant variables have been included in the model. Much of
our concern in the remainder of this chapter as well as in Chapter 6 is
to come to terms with this assumption. How do we know whether our
model is adequate inasmuch as it includes the main variables? As we shall
see, there is no easy answer to this. At this juncture, suffice it to say that
if our model is reasonably adequate, our error terms should behave as
noise (i.e. a normally distributed random variable with mean zero and a
constant variance ). If, however, the residuals of the estimated model
show signs that something has been left out, we should conclude that our
model is misspecified, although we may be unaware of the exact nature
of the problem. In fact, partial regression is built on this principle inas-
much as it uses the residuals of a regression which include explanatory
variables deemed relevant to check whether another variable adds
anything further in terms of making a significant contribution towards
explaining the variation in Y.

5.6 THE t-TEST IN MULTIPLE REGRESSION


Statistical inference through hypothesis testing becomes definitely more
interesting, varied and challenging as we move from simple to multiple
regression. The presence of several explanatory variables in the regres-
sion model invites us to explore more complex hypotheses. As in simple
regression, we may want to ask questions about the specific values a
coefficient can take. But in the case of multiple regression we are also
interested to see whether definite relations exists between two or more
regression coefficients. This section deals with the use of the t-test in
multiple regression analysis. The next chapter deals with the more versa-
tile F-test, which allows us to test for a whole range of linear restrictions
among regression coefficients. But this <loes not mean that the t-test is
of little use. This section shows that the t-test is a useful research tool in
applied analysis.
Partial regression 193
The t-statistic and the coefficient of partial correlation
The generalization of the use of the t-test to multiple regression analysis
is quite straightfarward. Given the assumptions of the linear model and
adding the assumption that the error terms are normally distributed, it
can be shown that the least squares estimators b1, b2 ••. bk are also
normally distributed. Analogous to the results obtained in simple regres-
sion it can then be shown that the variables
b - /3·
tj = ~; j = 1,2,3, ... , k (5.35)

each has a t-distribution with (n - k) degrees of freedom, where n is the


sample size and k the number of regression coefficients in the model.
The calculated t-statistics which are generally included in any software
package on regression or econometrics farmally tests the hypothesis H 0 :
13i = O, and hence involve computing the respective values of b/se(b) far
each regressor Xi" This test is used to see whether a single variable can
be dropped from the equation. In a sense, therefare, this test checks
whether or not a given variable contributes to explaining the variation in
the dependent variable.
This suggests that a relation should exist between a regressor's t-value,
thus calculated, and its partial coefficient of determination as defined in
section 5.4. This relation, stated without proof, is:
t-2
r 2 - J
Yj - t/ + (n - k) (5.36)

or
t-2
r 2 - J
Yj - t/ + degrees of freedom (5.37)
where ryi is the partial coefficient of correlation of regressor ~ with the
dependent variable Y, and ti is its calculated t-value under the hypothesis
H 0 : 13i = O. Consequently, it is always possible to calculate the partial
correlation coefficients corresponding to each regressor from the results
of a multiple regression analysis.
The implication of (5.36) (or (5.37)) is that the multiple regression line
contains all relevant infarmation about the partial regressions far each
of its explanatory variables. The partial regression coefficient equals
the corresponding slope coefficient of the multiple regression line and
the coefficient of partial correlation can be obtained directly from the
calculated t-statistic of the corresponding slope coefficient in the multiple
regression. There is no need, therefare, to sweep out the other regressors
in arder to arrive ata partial regression. However, the concept of sweeping
out is important since it teaches us how a multiple regression coefficient
is arrived at by controlling far the linear influence of the other regressors
194 Econometrics for developing countries
also included in the model. Moreover, to construct a partial regression
plot we need to be familiar with the method of partial regression. Let us
now illustrate the various uses of the t-test in multiple regression. Many
researchers only use the t-test to check the hypothesis whether a partic-
ular regressor has coefficient zero and, hence, should be dropped from
the equation. Obviously, this is an important application of the t-test.
But the t-statistic is more versatile in its use. For this reason, let us illus-
trate varied applications of the t-test using our example of section 5.2
concerning the demand for manufactured consumer goods in India.

Testing for a unitary elasticity


Starting with model specification (5.5), let us test whether we can simplify
the model by assuming a unitary income elasticity of the demand for
manufactured goods. Hence, our null hypothesis is:
Hº : 132 = 1 ; H1 : 132 #- 1 (5.38)
consequently, the resulting t-value is:

t = 0.959 - 1 = -0.63 (5.39)


0.0648
At 5 per cent level of significance, the critica! value of the t-distribution
with 18 ( = 21 - 3) degrees of freedom is t0 .05 = -2.l. The absolute
value of the calculated t-statistic is well below that for the critical value.
Consequently, the data allow us to maintain the null hypothesis of a
unitary elasticity.
We can now simplify the regression by imposing this unit elasticity on
the model. How do we do this? The general specification of the model is
given by equation (5.5). Imposing the assumption that 13 2 = 1, our model
beco mes:

log M = {31 + log (;,) + {33 log (-;:,) + E (5.40)

which, after rearranging terms, gives us the following specification:

log M - log (;,) = {31 + {33 log (-;:,) +e (5.41)

which reduces our model to a simple regression model. We can re-


estimate the model with least squares, obtaining the following results:

[ log M - log (I__)] = 2.82 - 0.406 log (~) (5.42)


p rn (0.01) (0.054) rn

R2 = 0.75, TSS = 0.1155; RSS = 0.0294


Partía! regression 195
This equation may look quite cumbersome, but if you take anti-loga-
rithms of both sides, you will obtain a neat average summary equation,
as follows:
p mM = e2.s2 (~)--0.4l
Y pm

= 16.8 (~r0.41 (5.43)

which tells us that the share of manufactured goods consumption in


income is inversely related to the relative price of food.
You may have noticed that the R 2 in regression (5.42) is much lower than
the R 2 obtained in regression (5.10). This should not give you cause for
concern, since the R 2s are not comparable as the regressions do not feature
the same dependent variable. However, the residual sums of squares of both
regression can be compared directly. Note that the RSS hardly increased as
a result of imposing the assumption that the income variable has a unitary
elasticity. We shall see in the next chapter that a comparison of the RSS from
restricted and unrestricted regressions is the basis for the F-test.

Exercise 5.4
Using model specification (5.3) which you estimated in exercise 5.2, test
the following hypotheses:
1 H 0: (3 2 = 1;
2 H 0: ¡3 3 =O;
3 Ho: ¡34 = -l.
In each case, specify clearly what you consider to be the relevant alternative
hypothesis. Explain the economic meaning of each of these hypotheses.

Testing for a linear combination between regression coefficients


Our final equation (5.43), which depicts the interplay between the demand
for manufactured goods and the relative price of food, was arrived at by
simplifying model specification (5.5) which, however, is itself a restricted
version of the broader model (5.2). So far we have assumed that there is
no money illusion; an assumption which allowed us to move from a four-
variable to a three-variable case. But are the data really comfortable with
this assumption? We should test whether the assumption:
(5.44)
is in fact data-admissible. To test this assumption with a t-test, we need to
reparameterise the coefficients of our general model as given in equation
(5.2): that is, we need to rearrange the terms in equation (5.2) in such a way
196 Econometrics for developing countries
that the sum of the slope coefficients in the original specification now
features as a single slope coefficient. To do this, add and subtract the term
(~ 2 + ~ 3 )log P m to the left-hand side of equation 5.2, leaving the overall
expression unchanged, as follows:
LogM = ~1 + ~ 2 logY + ~ 3 LogP1 + (~ 2 + ~ 3 + ~ 4 )1ogPm
- ~2logPm - ~3logPm +E
and, collecting all terms with the same coefficients, we get:

log M = /31 + /32 log (;:,) + {33 log (~) + (/32 + {33 + {34) log p m + E
(5.45)
which yields a specification which now features income deftated by indus-
trial prices, the relative price of food, and the price of manufactured goods,
all in logarithms, as explanatory variables. Importantly, the slope coeffi-
cient of the latter variable equals the sum of the coefficients ~ 2 , ~ 3 and
~ 4 • This enables us to test formally the homogeneity condition based on
the assumption of 'no money illusion'.

Exercise 5.5
Using specification 5.45, formally test H 0 : ~2 + ~3 + ~4 = O.

If you <lid the test, you will find that the data reject the null hypothesis. This
result may come as a surprise since the assumption that no money illusion
prevails appears quite reasonable in the light of demand theory. But, as
explained by Krishnaji (1992), our data are in fact highly aggregated time
series. As such, serious problems may emerge. For example, the effect of
rising average income per capita on the demand for manufactured goods
depends on how the increase in average income is distributed across house-
holds. If, as Krishnaji argues, rising average income goes hand in hand with
a worsening income distribution, the growth of the demand for manufac-
tured goods may conceivably be depressed accordingly. Similarly, a rise in
the price index <loes not mean that all prices go up at the same rate.
Differential price rises can have different consequences for incomes
and demand patterns. With aggregate data, therefore, the homogeneity
assumption is not as straightforward as it appears at first.
In exercise 5.4, you tested the hypothesis H 0: ~ 2 = 1 in model specifica-
tion (5.2) and found that the data reject the null hypothesis that the income
variable has an unitary elasticity. In other words, the demand for manu-
factured goods grows less than proportionally with income. Krishnaji's
explanation for this low income elasticity is that the worsening distribution
of income limits the expansion of the home market for manufactured goods.
But in fact Krishnaji <lid not use a double-log specification to estimate
Partial regression 197
the demand curve. Instead, in his specificatíon he tried explicitly to take
account of the dampening effect of rising incomes with worsening distrib-
ution on the demand for manufactured goods. He modelled the demand
for manufactured goods as a linear function of the price variables, P1 and
p m' and a quadratic function of income.

Testing for a quadratic term


Why this quadratic function? To see the point, it is instructive to start
with a linear specification in all variables:
(5.46)
Now, Krishnaji's argument is that, given the pattern of income growth
and its worsening distribution during the period concerned, the coefficient
of income cannot be assumed to be constant. Instead, he argued that this
coefficient is likely to fall as income per capita grows due to a worsening
income distribution. This requires that we model this coefficient as a func-
tion of the income variable itself. The simplest way to do this is by
assuming a linear function as follows:
(5.47)
Substitutíng equation (5.47) into equation (5.46) yields a quadratic term
income in the specificatíon of the demand for manufactured goods, as
follows:
M = 131 + l32P¡ + ¡33Pm + ¡34Y + 13sY2 + E (5.48)
which gives us Krishnaji's chosen specification.

Exercise 5.6
Using model specification (5.48):
1 Estímate the regression coefficients of the model.
2 Estímate the partíal regression of M and Y2.
3 Construct the partial regression plot of M and Y2.
4 Estímate the partial regression of M and Y.
5 Construct the partíal regression plot of M and Y.
6 Formally test the hypothesis Ha: 13 5 = O.
7 Formally test the hypothesis Ha: 13 3 = O.
8 Explain the economic significance of each of these hypotheses.
9 What does each partial regression plot (respectively, M on Y and M
on Y2) tell you?

This sectíon illustrates various uses of the t-test in multiple regression.


We have shown that the t-test can be used as a versatile instrument not
only to test hypotheses pertaining to one coefficient, but also those which
198 Econometrics for developing countries
involve linear combinations between a subset of coefficients. In the next
chapter we show the use of the F-test, which is even more flexible to test
for linear restrictions on regression coefficients - a characteristic which
makes it highly useful in the context of general to specific modelling.

5.7 FRAGILITY ANALYSIS: MAKING SENSE OF


REGRESSION COEFFICIENTS
An important conclusion from this chapter is that the estimate of the
effect of one variable (say, X¡) on Y, depends on the range of other
explanatory variables taken into consideration. Only in the case where X¡
is orthogonal to all other regressors in the model can we safely assess the
influence of X¡ on Y independently of the exact specification used in esti-
mating the coefficients. But orthogonality of regressors is rather rare in
applied socioeconomic analysis where we work with non-experimental
data. The special problem caused by multicollinearity is therefore largely
unavoidable with this type of data.
The problem is that collinear data provide 'weak evidence' (Leamer,
1978: 171) on the specific contribution of one or more individual regres-
sors. Typically, we end up with a very untidy regression: the R 2 may be
high but many coefficients of variables we believe to be important have
the wrong sign, take on an implausible value, or end up being statistically
insignificant due to their large standard errors. Furthermore, the subse-
quent exclusion or inclusion of a subset of regressors to the model can
often radically alter the estimated values of the regression coefficients in
the model. Hence, applied data analysis is often a messy business.
It is useful, therefore, to be aware how sensitive regression coefficients
can be across neighbouring specifications which involve different combi-
nations of collinear regressors. Much can be learned from looking carefully
at the behaviour of regression coefficients across alternative specifications.
To see this, let us look at a five-variable regression which is riddled with
multicollinearity.

Comparing regression coefficients across neighbouring specifications


Our interest is to investigate the socioeconomic determinants of fertility
levels across nations. Our dependent variable, therefore, is the total fertil-
ity rate, TFR. The total fertility rate is a demographic measure of the aver-
age number of children born to a woman, using age-specific fertility rates
for a given year. This measure, unlike the crude birth rate, is independent
of the age distribution of the population and hence lends itself better for
cross-country comparisons. To explain the variation in the TFR, we con-
sider three sets of regressors. The first, a single regressor, is an index of the
strength of family planning programmes, FP. We would indeed expect that
fertility depends, in part, on differences in family planning efforts across
Partial regression 199
countries. But, second, it is also reasonable to assume that fertility also
depends on the general socioeconomic context in which people are born,
live, work and die. It is difficult to find appropriate proxy variables to cap-
ture this social context. In our simple model, we shall use the logarithm
of gross national income per capita, log( GNPc), and child mortality, CM
(re-scaled per 100, instead of the usual 1,000, live births), as indicators of,
respectively, the income and the health of nations. Third, and finally, fer-
tility will obviously depend on the position of women in society - in
particular, the extent to which they can control their own fertility. This is
not easy to measure. We shall use female literacy, FL, as a very rough indi-
cator of the position of women in society (see, however, Sen, 1993, who
uses age at marriage as an alternative proxy to denote women's autonomy
in society).
Our regression model, including all five variables becomes:
(5.49)
subject to the usual assumptions of classical normal linear regression. To
estímate this broader model or any models which involve possible subsets
of these regressors, we use the cross-section data on 64 countries given
in the data file FERTILITY.

Specification bounds on regression coefficients


How sensitive are regression coefficients to alternative model specifica-
tions? In our example, we have four regressors. Let us first see what
happens to their coefficients if we combine them in different ways across
a range of specifications. How many different specifications are possible
with four regressors? Box 5.5 gives sorne simple rules on how to deter-
mine the number of various combinations of possible specifications to be
used in a sensitivity analysis.
In this example, we have four regressors and, hence, 15 (= 24 - 1)
possible regressions. Table 5.2 records all possible regressions with these
four variables. Befare discussing the results, it is useful to take a closer look
at the structure of the table. Three aspects are worth noting. First, the alter-
native regressions are grouped with respect to the number of explanatory
variables included in the regression: startíng with all simple regressions and
ending with the five-variable case. Second, within each group we follow a
lexicographic arder. Setting up the table in this way prevents us overlook-
ing one or other combination of regressors. Third, each regressor features
eight ( = 23) times in the set of alternative regressions.
If we look at the regression results listed in Table 5.1, we note that
there is quite considerable variation in regression coefficients across
different specifications. You can see this by running your eye down each
column, showing how the regression coefficient of a particular regressor
varies across the eight specifications in which this regressor is included.
200 Econometrics far developing countries
Table 5.1 Ali regressions with four regressors: the determinants of the total
fertility rate (TFR)
Dependent FP: family LG: lag GNP FL: female CM: child R2
Variable: TFR planning per capita literacy mortality
1 -0.042 0.57
2 -0.568 0.16
3 -0.036 0.39
4 0.133 0.45
5 -0.040 -0.453 0.67
6 -0.033 -0.022 0.69
7 -0.031 0.076 0.67
8 -0.038 -0.035 0.39
9 0.237 0.157 0.46
10 -0.013 0.096 0.47
11 -0.035 -0.234 -0.016 0.70
12 -0.034 -0.228 0.047 0.68
13 -0.031 -0.015 0.033 0.69
14 0.255 -0.014 0.119 0.48
15 -0.034 -0.212 -0.015 0.008 0.70

Box 5.5 Counting the number of alternative specifications

1 If there are (k - 1) regressors, the number of alternative regres-


sions we can run equals 2<k-I) - l. Why? Each regressor is either
included or not in any specific specification: hence, there are two
possibilities here. But we have (k - 1) regressors, each of which
is either included or not in any given specification, and, hence,
to arrive at all possible ways of combining these regressors in
alternative specifications we multiply the number of possibilities
of the inclusion/exclusion for each regressor (k - 1) times since
there are (k - 1) regressors. We thus arrive at 2<k-I) possibilities
in total. But this includes the case where none of the regressors
are included in the model. Subtracting this uninteresting case,
we arrive at 2<k-I) - 1 alternative specifications.
2 In practice, however, we do not always want to investigate all
possible alternative specifications given the number of regressors.
Often we are confident that sorne key regressors should always
feature in our regressions but we doubt the inclusion or exclu-
sion of the remaining candidate regressors. Hence, if there are
(k - 1) regressors of which k' are considered doubtful, we need
to consider 2k' alternative specifications. Note that, in this case,
the specifications which do not include the doubtful regressors
is also important: this is why there are 2k' cases and not 2k' - l.
Partial regression 201
Table 5.2 Minimum and maximum bounds of estimated regression coefficient
Bounds FP LG FL CM
Maximum -0.031 0.255 -0.013 0.157
Minimum -0.042 -D.568 -0.036 0.001

Clearly, sorne regressors have more stable coefficients across alternative


specifications than others. A quick summary can be obtained by listing
the mínimum and maximum bounds for each coefficient across its alter-
native specifications, as shown in Table 5.2.
Alternatively, we can also use techniques of exploratory data analysis
to look at the variation of the coefficients across specifications. Of partic-
ular relevance here is the use of single and comparative box plots. These
box plots reveal at a glance the average level of the point estimates, their
spread, skewness as well as possible outliers. Figure 5.6 shows the compar-
ative box plot for our coefficient estimates. Note that, to construct this
plot, we standardised each coefficient estímate by dividing it by the
absolute value of its mean across the different specifications which include
that coefficient.
The plot shows that the estimates of the family planning variable are
remarkably stable across specifications. The estimates fluctuate within very
narrow bounds. In contrast, the coefficient of the income variable is much

2 -
-~
¡--

_l__
o

1 1
-1
---'--- 1 1
_L_
-2 -

-3 -

-~

-4 -
s1a1a™
Figure 5.6 Variation in regression coefficients across specifications
Note: From left to right: FP; log(GNPc); FL; and CM
202 Econometrics far developing countries
more fragile with respect to alternative specifications, occasionally even
producing the wrong sign. The coefficients of child mortality and female
literacy vary significantly across specifications without, however, producing
the wrong sign in any of them. Obviously, there is a fair amount of multi-
collinearity among regressors. To see this, take a good look at the last
column in Table 5.1. None of the multiple regressions have an R 2 which
is anywhere near the sum of R 2s of the simple regressions featuring their
corresponding regressors. The table also reveals sorne cases of the pres-
ence of superfluous variables. In specification 15, for example, the
coefficient of child mortality dwindled almost to zero and, hence, drop-
ping this variable from the regression hardly affects the regression
coefficients of the other regressors, as can be seen from specification 11.
Similarly, in specification 8, the coefficient of the income variable is excep-
tionally close to zero; dropping it from the regression hardly affects
the regression coefficient of the other regressor, FL, as can be seen from
specification 3.

Conditional bounds on regression coefficients


Up to this point we have looked at the maximum and minimum bounds for
each coefficient across all possible combinations of regressors in the model.
This procedure can be useful to pick out coefficients which are remarkably
stable across various specifications, but it may at times be misleading.
Indeed, sorne regressors will only have meaningful coefficients if they fea-
ture in the model along with other regressors. In a demand equation, for
example, the inclusion of a price variable is unlikely to give good results
if the income variable is not included in the model. Hence, in most appli-
cations we are not interested in the maximum and minimum bounds of
regression coefficients across all possible specifications, but instead we are
likely to narrow down the range to a subset of reasonable specifications.
Obviously, it is always important to explain clearly what you consider to
be the relevant set of specifications to be taken into consideration.
Let us try this idea with our example. First, a closer look at the table
with alternative specifications shows that, if we include either female
literacy or child mortality in the regression, the bounds of the family plan-
ning variable become even more proximate to each other: the coefficient
values now lie within the range -0.31 to -0.35! The fact that this range
narrows down once we include any one of these social indicators should
not surprise us: family planning is not isolated from the general socio-
economic context and from the position of women in society. It is sensible,
therefore, to remove the influence of general social conditions in order
to assess the separate effects of family planning.
Similarly, it is clear that the family planning variable should not be
discarded from a serious analysis and, hence, we should only consider the
conditional bounds of the coefficients of the other regressors in those
Partial regression 203
Table 5.3 Conditional bounds of coefficients of FL and CM: FP included
Bounds Fema/e literacy Child mortality Income per capita
Maximum -0.015 --0.076 -0.212
Minimum --0.022 -0.008 -0.453

regressions which include the family planning variable. Table 5.3 gives us
the conditional bounds of the coefficients of FL and of CM in regressions
which include the family planning variable. As you can see, the bounds
have now become considerably closer. Undoubtedly, the female literacy
variable is less fragile in the sense that its point estímate is much less
sensitive to the range of specifications under consideration. The income
variable retains a relatively large fluctuation in its coefficient, but the bond
is also narrowed and all values have the 'right sign'.

The question of competing proxy variables


A closer look at Table 5.3 reveals that the joint presence of FL and CM
in the model invariably affects the coefficient estímate and the statistical
significance of either one of these variables. It appears, therefore, that
both variables compete for attention as rival proxies for a similar under-
lying variable. Yet clearly the female literacy variable performs best.
A plausible explanation is the following. Both variables are in fact social
indicators and, as such, are proxies for general underlying social condi-
tions. But the female literacy variable captures something in addition to
this: it also functions as a rough proxy for women's autonomy or status
in society. In fact, you could convincingly argue that female literacy goes
more to the heart of the matter since it is indicative of conditions which
may be equally important in explaining child mortality as well as fertility
behaviour. It is not surprising, therefore, that female literacy proves to
be a more robust regressor in the model.
In view of our discussion above, it seems sensible to conclude that model
specification 6 (in Table 5.1) which features the regression with two
explanatory variables - family planning effort and female literacy - is
probably the most suitable for our purposes, given the initial specification
from which we started. Little or nothing is gained by adding the other
two explanatory variables.

Lessons
This example shows that sensitivity analysis can be a great help in guard-
ing against making fragile inferences. Looking carefully at the behaviour
of regression coefficients across altemative specifications often gives us
useful insights. When one or more coefficients prove to be highly robust
with respect to altemative specifications we may feel more confident about
204 Econometrics for developing countries
the inferences we make about the influence exerted by such variables. But
this <loes not mean that we should look at other variables with suspicion.
Indeed, quite often one or another coefficient of a regressor only becomes
reasonably stable once one or more other regressors are already included
in the model. In fact, in any regression there are usually sorne variables we
feel confident should always feature in the regression model. However, it
is always important to check how stable such coefficients are once we
include or exclude other variables which we consider to be doubtful, super-
fluous or of minar importance. Finally, our example also taught us that
sorne of the regressors we take into consideration when searching for an
appropriate specification may in fact be different proxy variables of a
deeper variable which may be hard to measure directly. The joint presence
of such related proxy variables often renders their coefficients unstable or
insignificant, or both. Sensitivity analysis may help to spot such problems
and help us to select the proxy which seems most appropriate.

Other things being equal?


Can we interpret a regression coefficient as the marginal impact of its
regressor on the dependent variable, other things being equal, when in
fact we know other things were not at all equal? This is a tricky question
which requires careful reflection. One thing is certain: it is important to
avoid giving a mechanical interpretation to any regression coefficient.
Given a model such as:
(5.50)
it has become far too common to interpret the coefficient of X 2 as the mar-
ginal impact of X 2 on Y, other things being equal (i.e. holding X 3 constant).
This interpretation is most definitely not always valid (Mosteller and Tukey,
1977: 300). Consider, for example, the following regression model:
(5.51)
which expresses a quadratic trend line. It is clearly wrong to interpret 13 2
as the marginal impact of time on Y holding t 2 constant. In this case the
error we make is obvious since the same variable appears twice in the
regression line: once linearly and once as a quadratic term. But often we
find regressors jointly in a model which are either related to one another
or serve as alternative proxies for the same deeper variable we cannot
hope to measure accurately. For example, can we really argue that the
coefficient of female literacy in our model gives us the effect of a unit
change in female literacy holding other things (including child mortality)
constant? In fact, as we argued above, female literacy may well be a deter-
minant of both child mortality and fertility behaviour. Regression allows
us to come to grips with the separate influence of a regressor on the
dependent variable over and above the combined influence exerted by
Partía! regression 205
other regressors. But this does not necessarily mean that this reftects the
effect of this regressor on the dependent variable, holding other things
constan t.
Hence, the interpretation of regression coefficients always requires
careful reftection. In sorne cases it may be plausible to interpret the coef-
ficient as if other things were held constant. But we should never jump
too quickly to conclusions. For example, we may rashly assume that, other
things being equal, a proportional increase in income per capita and in
prices will leave demand for manufactured goods unchanged. But with
real data, as we have seen, changes in the distribution of income may well
render our estimates of the coefficients in the model contingent on the
actually prevailing conditions where things are never equal. The elastici-
ties we measure in such a case are rather complex and composite measures
which reftect that other things are changing all the time, for example, the
distribution of incomes.
The mechanical interpretation of regression coefficients, therefore, can
lead us astray in many ways, particularly if we intend to base policy conclu-
sions upon the premise that a single coefficient of a regression measures
the impact of its variable on the dependent variable, other things being
equal. Other things are never equal, and often the pattern of covariation
between regressors is part of the fabric which structures the phenomenon
in question. Overlooking this fact can lead to simplistic conclusions being
drawn from the application of sophisticated statistical techniques.

Exercise 5. 7
Model specification (5.2) is, in fact, a restricted version of a more elabo-
rate model which include, apart from the income variable, Y, the price
variables, P1 (the price of cereals) and Pm' two more price variables:
namely, P 01, a price index of other food products, and P8 , a price index
of consumer services. These data are in the data file KRISHNAJI.
Including the last two variables into the double-log model specification
yields a six-variable regression.
1 Construct a table with the results of all possible regressions which at
least include the income variable (why?)
2 Construct comparative box plots of the variation in the slope coeffi-
cient of each regressor in the model.
3 Judging from your table, check whether there is much evidence of
multicollinearity.
4 Check whether any variables in any of the specifications appear super-
ftuous.
5 How robust is the income elasticity across alternative specifications?
6 In your opinion, which price variables appear to be most relevant in
the model?
206 Econometrics for developing countries
5.8 SUMMARY OF MAIN POINTS
1 Multiple regression is nothing more than a hierarchy of simple regres-
sions involving regressions between the dependent variable and each
of the regressors, auxiliary regressions between regressors, and regres-
sions featuring as its variables residuals of prior regressions.
2 The regression coefficient of an explanatory variable in a multiple
regression equals that of the partial regression arrived at by sweeping
out the linear influence of the other regressors included in the model
from both the dependent variable and the explanatory variable in ques-
tion. This procedure yields a partial regression which allows us to
control for the influence of other variables while investigating the rela-
tion between the dependent variable and an explanatory variable in a
context where other things are not equal.
3 The coefficient of partial correlation is the square root of the coeffi-
cient of determination of the partial regression. It is the coefficient of
correlation between the dependent variable and the added regressor
after sweeping out the influence of the other regressors in the model.
4 A partial regression plot (or added-variable plot) is a scatter plot
between two sets of residuals obtained by removing the linear influ-
ence of the other regressors from both the dependent variable and the
added regressor. A partial regression plot allows us to look at a
multiple regression coefficient by means of a two-dimensional scatter
plot. It is a powerful diagnostic tool to detect deviations from model
assumptions.
5 The extension of the least squares principie to multiple regression is
quite straightforward and yields mathematical properties which are
similar to those of simple regression. The main difference is that there
is now more than one explanatory variable, and with non-experimental
data, these variables often display a fair amount of multicollinearity.
Perfect collinearity implies that an exact linear relation exists between
the regressors (including the constant term), in which case linear
regression breaks down. Orthogonality of regressors implies that they
are uncorrelated with one another, a situation which is ideal for regres-
sion but seldom satisfied in practice.
6 Due to the presence of collinearity, the slope coefficient of a given
regressor in relation to the dependent variable depends on the other
regressors included in the model. Consequently, the slope coefficient
varíes with model specification as a result of the inclusion or exclu-
sion of other regressors. Only if all regressors are orthogonal on each
other will simple regression yield the same slope coefficients as
multiple regressions.
7 A superfluous variable in a regression is one which adds nothing to the
explained variation once the effect of other regressors have been taken
into account. Strictly speaking, its slope coefficient in the multiple
Partial regression 207
regression will be zero, but it may well be non-zero in subset regres-
sions drawn from this broader model. Dropping a superfiuous variable
from the regression does not alter the slope coefficients of the other
regressors.
8 Given the usual assumptions of classical linear regression jointly with
the added assumption that regressors are not exactly collinear, the
least squares estimators turn out the be BLUE. If we add the normality
assumption, the least squares estimators are also ML estimators and,
therefore, have minimum variances among all estimators. Given these
assumptions, the least square line is unbeatable as an estimator of the
population regression.
9 The t-statistic allows us not only to test hypotheses concerning indi-
vidual values of the coefficients, but also hypotheses which involve
linear combinations of coefficients provided we can suitably repara-
meterise the model.
10 It is a useful exercise to investigate the sensitivity of regression coeffi-
cients across plausible neighbouring specifications to check the fragility
of the inferences we make on the basis of any one specification with
respect to specification uncertainty as to which variables to include.
11 An important conclusion of this chapter is that we should not to readily
be led to assume that a slope coefficient in a multiple regression
measures the marginal impact of its regressor on the dependent vari-
able, other things being equal. With non-experimental data, other
things are never equal and hence our estimation and hypothesis testing
always takes place in a context where the covariation between regres-
sors is part of the picture. Regression only allows us to remove the
linear infiuence of other regressors when looking at the bivariate rela-
tion between Y and a particular regressor, but this is by no means the
same as holding the other regressors constant. Whether or not such
an inference can be made requires careful refiection, not a leap of faith
or blind trust.
6 Model selection and misspecification
in multiple regression

6.1 INTRODUCTION
In the last chapter we saw that any particular slope coefficient of a regres-
sor generally depends on the other regressors included in the model. Hence,
even if we are interested in the impact of only one of the regressors, it is
important that we include all relevant variables in the regression equation.
In this chapter we take this point further and show that if we omit a rele-
vant variable from the model, least squares will no longer give us unbiased
estimators of the coefficients of the population regression. The problem
is misspecification due to the omitted variable bias. That is, unless all
relevant variables are included in the regression equation, then none of the
estimated parameters will be unbiased (except in a special case shown
below). However, this result should not lead to a strategy of including every
conceivable variable one can think off in a regression since, with collinear
data, regression results are likely to yield foggy messages due to infiated
standard errors of the coefficients. Variable selection in model specification
is, therefore, a challenging task in applied research and has serious conse-
quences for the validity of the inferences we make from our data.
This chapter deals with hypothesis testing in the context of model selec-
tion and hence introduces you to sorne of the basic principies of general to
specific modelling. In section 6.2, we begin this chapter with an example of
misspecification: Griffin's well-known argument that aid displaces savings.
We show how, according to his own theory, Griffin's equation was mis-
specified. Next, in section 6.3, we examine the theory behind omitted vari-
able bias as well as the implications of omitting relevant variables or adding
irrelevant variables for the standard errors of regression coefficients, and
relate this discussion to our examination of Griffin's model. Excluding vari-
ables from a model is a form of restriction being imposed on a more gen-
eral model which includes these variables. This point is explained in section
6.4, where the use of F- and t-tests in specification searches is explored.
The F-test can be used to test any linear restrictions we place on a
model: not just testing for the exclusion of certain variables (zero restric-
tions ), but also testing for particular linear relations between regression
Model selection and misspecification 209
coefficients (non-zero restrictions) and for pooling data from different
samples (including different time periods). These issues are discussed in
sections 6.5 and 6.6. Even if it is not valid to pool the data in a partic-
ular sample, it may still be possible to estimate a single equation by the
use of dummy variables to allow sorne variation in coefficients between
sub-samples, this application is explored in section 6.7. Section 6.8
summarises the main points from this chapter.

6.2 GRIFFIN'S AID VERSUS SAVINGS MODEL:


THE OMITTED VARIABLE BIAS
During the 1960s there was a very optimistic view of aid's role in economic
growth. Aid would supplement domestic savings, allowing higher invest-
ment which (through a Harrod-Domar equation) would result in higher
growth (see Box 6.1). By the end of the decade a greater pessimism was
developing and doubts began to be expressed about aid's effectiveness in
increasing growth. Theoretical backing for these doubts was given by
Griffin (1970; see also Griffin and Enos, 1971) who argued that aid would
not simply supplement domestic savings, as had been hitherto expected,
but rather displace at least sorne part of them. The rise in investment
resulting from higher aid would thus be rather less than the one-to-one
increment assumed in the optimistic 1960s growth models.
In a response to criticisms of his original paper, Griffin (1971) re-
presented his argument in the following model. Consumption is given by:
C = a + 13 (Y + A) (6.1)
where C is consumption, Y is income and A aid. Since savings (S) are
given by:
S=Y-C (6.2)
it follows that:
S = -a + (1 - 13) Y - 13A (6.3)
demonstrating the negative relationship between aid and savings. Griffin
then, in his own words, 'suppresses' (i.e. ignores) the constant term and,
dividing through by Y, gets a model to be estimated of the form:

(6.4)

where his theory would suggest that 13 2 < O.


In his 1970 paper, Griffin presented four estimates of equation (6.4),
in which the slope coefficient ranged from -0.67 to -0.84. We re-estimated
the equation using 1987 data for 66 developing countries. These data, which
are used throughout this chapter, are given in the file AIDSAV on the
diskette. Estimation of equation (6.4) gives (standard errors in parentheses):
210 Econometrics far developing countries

Box 6.1 Aid in the Harrod-Domer Model


The Harrod-Domer equation is an identity which relates the rate
of growth of output (Y) to the level of investment:
1 I
g=- -
k y
where !/Y is the investment rate and k is the incremental
capital-output ratio (ICOR). This identity becomes a theory of
growth if we assume capital to be the critical constraint on output,
i.e. higher investment is a necessary and sufficient condition for
increased output. Investment is given by the accounting identity:
l=S+A
where S is gross domestic savings and A foreign savings (i.e. capital
inftows, which are largely aid for many developing countries). This
identity may be expressed as a percentage of GDP (Y):
I S A
-
y =-
y + y-
It is evident that higher aid (A/Y) will increase !/Y and hence growth.
Suppose initially there is no aid and the domestic savings rate is 9
per cent. An ICOR of 3 will translate these savings to a growth rate
of 3 per cent. Suppose now that the country receives aid equivalent
to 6 per cent of GDP. The investment rate rises from 9 to 15 per
cent and growth increases to 5 per cent.
Griffin contested this view, arguing that aid would displace savings
so that the investment rate would rise by less than the increase in
A/Y.

1\
s 19.14 - 0.87 -
A
R2 0.33
y
= (6.5)
(2.06) (0.16) y RSS = 10,358

The coefficient of -0.87, which is not very different from Griffin's results,
is significant at the 1 per cent level and appears to confirm the argument
that aid will displace savings. But we saw above that the estimated model
should not be equation (6.4) at all, since to get it Griffin ignored the inter-
cept term in equation (6.3). As the estimated equation is divided through
by income, the 'true model' should also include the reciproca! of income
on the right-hand side, that is:

(6.6)
Model selection and misspecification 211
Estimation of equation (6.6) yields:
A
s A 1
- = 20.90 - 0.40 - - 17375 - R 2 = 0.57 (6.7)
y (l.69) (0.15) y (2923) y RSS = 6,637
The magnitude of the negative relationship between aid and savings is
halved by estimating the correct equation derived from Griffin's model,
vividly illustrating the point made in the last chapter that the value of an
estimated coefficient depends on the other regressors included in the
model. Befare we go on to look at the theory behind this omitted vari-
able bias, it is useful to pause and consider what we have done above.

Encompassing in empirical research


There are many unresolved debates in the social sciences, and development
research is no exception. Empirical research often seems to further con-
fuse, rather than resolve, these debates. The encompassing approach seeks
to isolate one such source of confusion, by first analysing if different results
are simply the results of different data sets. Many issues have been analysed
using a range of different data sets. Therefore different conclusions may be
put down to different data. In the case of the aid-savings debate, Riddell
wrote that the: 'negative relationship between capital inflow and domestic
savings was confirmed by Landau for 18 Latin American countries.
However, Gupta's 1970 results for 31 countries indicate the foreign capital
inflow has had little effect on domestic savings ratios' (Riddell, 1987: 118).
But if the researchers are using both different data and different models
there really is no basis of comparison for their results. For sorne problems
- say those for a specific country - it would be helpful if the same data set
were used, and sorne authors publish their data to facilitate replication of
their results. This practice is a very helpful one.
In the case of Griffin's analysis his data set was not available to us. But,
as a next best thing, we estimated his model using our data. We found
that we reproduced (or replicated) his result fairly closely. Such replica-
tion is a very important stage of empirical research: it means that
differences between our conclusions and those of previous researchers do
not arise out of differences in the data. Therefore, it would be wrong to
say that 'using 1960s data Griffin found a displacement effect of up to
-0.82, whereas with more recent data a smaller effect of only -0.40 was
found'. This conclusion might be valid if we had not first replicated
Griffin's result. Our estimated slope coefficient of -0.40, however, is not
a product of the data, but of our model specification.
Our example shows that inadvertently dropping a relevant variable from
the model can lead to significant changes in our coefficient estimates. In
section 6.4, we shall try to test whether or not Griffin's equation (6.4)
is an invalid restriction of the correct model as specified in equation
212 Econometrics far developing countries
(6.6). But before we do so, we need to equip ourselves with sorne more
theoretical tools concerning the omitted variable bias. This we do first in
the next section.

Exercise 6.1
Use the data in the file MALTA (Maltese exports demand and supply)
to estimate the following demand equation for exports, X:
ln(XJ = ¡3 1 + ¡3 2 ln(WYJ + E1,;

where WY is a variable measuring (world) income: the income of Malta's


main trading partners.
1 How would you interpret the slope coefficient from this regression?
Next, estimate the equation:

ln(XJ = ¡3 1 + ¡32 ln(WYJ + ¡3 3 ln(PXJ + Ez;

where PX is an index of export prices.


2 How do these results alter your conclusions from the simple regression?

6.3 OMITTED VARIABLE BIAS: THE THEORY


What are the implications of erroneously omitting a relevant variable from
a model? To tackle this question, let us first use the three-variable case.
Suppose that the 'correct model,' also known as the data generation
process (DGP), is as follows:
Y; = ¡3 1 + ¡32X 2; + ¡33X 3; + E; (6.8)
but that, overlooking the importance of X 3 , we wrongly specify the model
as:
(6.9)
What then are the implications of omitting X 3 for our estimate of the
slope coefficient of X 2 ?
You will recall from Chapter 4 that the OLS estimator of ¡3 2 , bY2, in
the simple regression (6.9) is given by:
n

b =
L (X 2; - X)Y;
~i=~l_ _ _ __ (6.10)
Y2 n

L (X2; -X)
i=l
2

If model (6.9) were correct, by2 will be an unbiased estimator of ¡32 •


However, by assumption, we know the correct model is given by equation
(6.8). Substituting the correct expression for Y; into equation (6.10), yields:
Model selection and misspecification 213
n

2: (X2; - X) (f31 + f32X2; + f32X2; + e;)


b 2 -- ~;~~1~-------------
n (6.11)
2: (X
;~1
2; - X) 2

If we work out the product in the numerator, we shall see that the first
term disappears since the sum of deviations from the mean of X equals
zero. The second term will reduce to ~ 2 , as its numerator and denomi-
nator are equal.
Taking expectations of both sides of equation (6.11) then yields:
n

2: (X2; - X)X3;
E( b2) = f32 + {33 + _;~-~----- (6.12)
:¿ (X2; -X)2
;~1

since the X-variables are given and, hence, non-stochastic, and E(E;) =O.
Equation (6.12) shows that byz yields a biased estimator of ~ 2 • This bias
is the product of two terms. The first term, ~ 3 , is the population slope
coefficient which measures the impact of the wrongly omitted variable on
the regressand. The second term is the estimate, c2, of the slope coeffi-
cient of the following auxiliary regression:
Y3; = "'11 + -Y2X2; + E; (6.13)
Hence, equation (6.12) can also be rewritten as:
E(bz} = ~ 2 + ~ 3 c2 (6.14)
Equation (6.14) allows us to determine the direction of the bias. If
the relationship between the omitted variable and the regressand (X3 and
Y) and the correlation between the omitted and included variables
(X3 and X 2) have the same sign (i.e. both are positive or both negative),
then their product is positive and there is an upward bias. If the two
expressions have different signs, their product is negative and there is a
downward bias.

The exceptions: orthogonal regressors and irrelevant variables


This last expression of the bias allows us to see when omitting a variable
does not lead to any bias in the estimates of the slope coefficients of the
other regressors. More specifically, there will be no omitted variable bias
if, either (a) ~ 3 = O, which would indicate that X 3 is in fact an irrelevant
variable in the model, and hence, can conveniently be dropped (in which
case equation (6.9) gives us a correct specification), or (b) X 3 happens to
be uncorrelated (orthogonal) with X 2 in the sample and, hence, the slope
coefficient in the auxiliary regression of X 3 on X 2 is zero.
You will have noted the striking similarities of these formal results with
our discussion in Chapter 5 on the behaviour of regression coefficients
214 Econometrics far developing countries
across different specifications which involve the inclusion/exclusion of
other regressors. The case of orthogonality is exactly the same as before.
If regressors are orthogonal, no omitted variable bias will occur and, simi-
larly, simple regression estimates yield the same slope coefficients as
multiple regressions. This is an ideal situation which, unfortunately, is
seldom found when working with real data.
The case of the inclusion of irrelevant variables is similar to, but not
identical with, that of what we called superftuous variables. A variable X¡
is irrelevant if its slope coefficient of the population regression equals
zero: i.e. 13¡ = O. A variable is superftuous in a particular regression,
however, if its estimated (sample) slope coefficient equals zero: b¡ = O.
These are not equivalent statements. lt is perfectly possible that a rele-
vant variable yields a zero slope estimate in a particular sample, perhaps
due to excessive multicollinearity or unfortunate sampling. Similarly, the
fact that a variable is irrelevant <loes not mean that it may not produce
a non-zero (or even significant) slope estimate in a particular regression.
Obviously, under conditions of repeated sampling, an irrelevant variable
will produce coefficient estimates which, on average, equal zero.

Assessing the direction of the bias


Let us now return to Griffin's model. We found that our estimation of his
restricted model led to a downward pull on the estimate (from -0.40 to
-0.87). In fact, this is what we should expect to happen. Why is this? As
we have just seen, the direction of the bias depends on the sign of the
product of two terms: (a) the sign of the slope coefficient of the omitted
variable in the population regression of the correct model; and (b) the
sign of the correlation coefficient between the omitted and included
regressors (the correlation coefficient between two variables has the same
sign as the regression coefficient). Equation (6.3) shows that the first of
these should be negative. But what about the correlation between 1/Y and
A/Y? The omitted variable, 1/Y, will be small for two reasons: if a country
is either big or rich. Small countries are known, on average, to receive
disproportionately more aid than large ones. Also rich countries may
expect to receive less aid than poor ones. For both these reasons, as
income rises (so 1/Y falls) the aid ratio (A/Y) will also fall. The correla-
tion between l/Y and A/Y is therefore positive and the sign of the omitted
variable bias (the product of a negative and a positive) is negative.
Equation (6.12) can also be used to estimate the value of the bias. We
do not know 132 , but we estimated it to be -17 ,375, as shown in equation
(6.7). We also need to know the coefficient from regressing 1/Y on A/Y,
which yields a coefficient of 0.000026 (small, but significant: the t-statistic
is 4.98). The product of these two is -0.47: which is exactly equal to the
difference between the estimates of this slope coefficient obtained from
equations (6.5) and (6.7): respectively, -0.87 and -0.40
Model selection and misspecification 215
Exercise 6.2
Using Exercise 6.1, assess the direction of the bias obtained by estimating
the simple regression if, in fact, the three-variable model is the correct
specification. Is your theoretical analysis verified by your empirical results?

Omitted variable bias in the k-variable case


The extension of the omitted variable bias to the k-variable case is quite
straightforward. In the case where our estimated model includes (k - 2)
explanatory variables, X 2, ••• , Xk_ 1, but the correct model also includes
the relevant variable Xk, the omitted variable bias of the slope coefficient
of the included regressor X¡ is obtained as follows:
E(b¡) = f3¡ + ¡3k Ck¡; j = 2,3, ... , (k - 1) (6.15)
where ck¡ is the slope coefficient of X¡ in the auxiliary multiple regression
of Xk, the omitted variable, on all the regressors included in the model.
As you can see, equation (6.13) is a special case of this equation.
What happens if we omit more than one relevant variable from
the model? Suppose our estimated model included (k' - 1) regressors,
X 2 , ••• , Xk'' and omitted (k - k') regressors, Xk' + 1, •.. , Xk , from the
correct model. In this case, the omitted variable bias for the slope coef-
ficient of X¡ is given by:
k
E(b¡) = /3¡ + ~ (/3;C¡¡); j = 2,3, ... 'k'; l = (k' + 1), ... 'k (6.16)
l=k'+l

where c1¡ is the slope coefficient of X¡ in the auxiliary multiple regression


of X 1, an omitted variable, on all regressors included in the model.
Equation (6.15) gives us the most general case of the omitted variable
bias. Obviously, the more complex the situation in terms of the number
of variables included and excluded from the model, the more difficult it
becomes to make any a priori statements about the likely sign of the bias.
But as you can easily verify, the slope coefficient of X¡, a variable included
in the model, will not be affected by any omitted variables bias if (a) the
omitted variables are all orthogonal on X¡, or (b) if they all happen to be
irrelevant and, hence, can rightfully be omitted from the equation.

What about the standard errors of the regression coefficients?


We have seen that omitting one or more relevant variable from a regres-
sion renders the least squares estimators of the coefficients of the model
biased. Hence, the mean of the sampling distribution of a regression coef-
ficient will no longer equal the corresponding population parameter. But
what about the variance of this sampling distribution? How is this affected
by omitting a relevant variable from the model? In this case, the outcome
216 Econometrics far developing countries
is more ambiguous.
To see this, recall from Chapter 5 that the variance of the least squares
estimator of a slope coefficient of a multiple regression is given by:
a2
V(/3) = s._ · V!Fj ; J. = 2,3,. .. , k (6.17)
]]

Where V!Fj is the variance inftation factor of the slope coefficient of


regressor Xj, and S¡j is the sum of squared deviation of X¡ from its mean.
V!Fj = 1/(1 - R/), where R/ is the coefficient of determination of the
auxiliary regression of X¡ with all other regressors included in the model.
The variance of the estimator of the slope coefficient of a regressor X¡,
therefore, depends on the three factors: the error variance, the total sum
of squares of X¡, and its variance inftation factor.
Now, it can be shown that if we omit one or more relevant variables
from the model, the variance of the estimator of the slope coefficient of
regressor X¡ in the model which omits one or more relevant variables is
also given by equation (6.17) with the exception that its VIF¡ is now
derived from the auxiliary regression of X¡ with the smaller set of regres-
sors included in this simpler model. Now, since the simpler model features
less regressors than the larger model, it follows that the variance inftation
factor of the former will be less than (or at most equal to) that of the
latter. The reason is that the more variables we include in the model, the
greater the multicollinearity, and, hence, the greater the R 2 of the auxil-
iary regressions, unless the omitted variables happen to be orthogonal to
the regressors included in the simpler model. The crux of the matter is
that the simpler model which omits relevant variables produces biased
estimates but with smaller variances! Consequently, there appears to be
a trade-off between bias and precision.
But the story <loes not end here. Equation (6.17) features the variance
of the error term which in practice is unknown to us. To estimate the
standard errors of the regression coefficients we estimate this variance
using the residual sums of squares of the estimated model as shown in
equation (5.34) in the previous chapter; that is:
s2 = RSS/(n - k) (6.18)
But this estimate will differ depending upon the number of regressors
included in the model. To see this, let us see what happens to, respec-
tively, the numerator and to the denominator of s2 if we include omitted
variables. As to the numerator in (6.18), the RSS of the larger model will
generally be less than that of the model which omits relevant variables.
In particular, when we omit an important relevant variable, the gain in
terms of reducing the residual sums of squares obtained by including it
in the model is likely to be considerable. As to the denominator, given
the sample size, smaller models have more degrees of freedom. In sum,
Model selection and misspecification 217
whether the inclusion of omitted variables in the model leads to a reduc-
tion in s2, the estimated standard error of the regression, depends on
whether the RSS falls sufficiently to offset the loss in degrees of freedom.
In general, if our omitted variables play an important role in explaining
the variation in the dependent variable, we would expect the estimated
error variance to fall with the inclusion of the omitted variables in the
regression, unless the sample size is very small. This last point is obvious,
but often overlooked. There is not much point, for example, in estimating
a regression with five or more variables if we have only a handful of obser-
vations.
So what will happen to the estimated standard errors of regression coef-
ficients if we add omitted variables to the model? Our discussion above
shows that two opposite tendencies are at work: adding relevant variables
to the model generally reduces the estimated error variance, s2 (unless
the loss in degrees of freedom offsets the gain in the reduction in the
RSS), but increases the V!Fs, the variance inftation factors of the slope
coefficients due to multicollinearity. The net outcome cannot be stated a
priori. It differs from case to case. If the reduction in the estimated error
variance dominates the scene, the standard errors of the regression coef-
ficients will decrease with the introduction of the omitted variables in the
model, and vice versa if the variance inftation factor gains the upper hand.
What practica! lessons can we learn from this discussion on bias and
precision due to omitting relevant variables? First, omitting an important
variable which has a strong impact on the dependent variable produces a
bias as well as larger standard errors for the regression coefficients. In this
case, in practica! terms, there is no real trade-off: omitting the variable
worsens both bias and precision. For example, this generally happens when
the income variable in a demand equation is omitted: the resulting regres-
sion will not make much sense. Second, if we omit a relevant variable which
has a more limited impact on the dependent variable and which correlates
strongly with other regressors already included in the model, the trade-off
between bias and precision may well be very real. The reduction in bias is
likely to be bought at the price of less precision. For example, in a demand
equation, the inclusion of a minar price variable of a more distant substi-
tute may well make matters worse even if there are sound theoretical
reasons for the variable to play a role, although not a majar one. In such
cases, we often end up with a large R 2 and a large number of insignificant
coefficients (as was the case with the example in exercise 5.7). Finally, if we
only have a small sample at hand, we cannot hope to include all relevant
variables and still get reasonably precise results. The smaller the sample
size, the more the loss in degrees of freedom due to the inclusion of added
variables matters in terms of precision. Regressions with many variables
and few data points are unlikely to produce firm inferences.
218 Econometrics far developing countries
The problem of irrelevant variables
This analysis of the trade-off between bias and precision leads us to an
important but often neglected point: a regression is not a sack into which
we can throw any variable we can think of. Even with relevant variables
we often have to decide whether the reduction in bias is worth the loss in
precision. Not surprisingly, the problem gets worse if we take a lot of irrel-
evant variables on board as well. Hence, the bad news is that irrelevant
variables boost the variance inflation factors of regression coefficients
without yielding any gain in terms of reducing the estimated standard
error of the regression. The greater the collinearity between variables, the
greater the loss in precision will be. But the good news is that, as you would
expect, the inclusion of irrelevant variables does not affect the unbiased-
ness of our estimators (if the model is otherwise correctly specified).
A final point: often the problem is not so much that we attempt to
burden a regression with a number of irrelevant variables, but rather that
our stock of potential regressors includes one or more subsets of proxy
variables which, within each set, essentially seek to measure the same
thing. We carne across an example when discussing the determinant of
fertility levels across countries in Chapter 5. Our point is not that we
should not try out various proxies. Indeed, proxy variable specification
searches are important in many regression applications. Instead, our point
is that we should always keep in mind that sorne of the variables we select
for consideration are chosen to serve as proxies for sorne deeper under-
lying variables which we cannot measure. The danger is, however, that
once these variables enter a regression, we no longer see them as rival
proxies, but treat them as separate variables in their own right.

Exercise 6.3
This exercise requires that you generate a set of artificial data to check
on the implications of adding irrelevant variables to a model. Start with
the assumption that the correct specification involves a simple regression
model: Y= 10 + 0.6 X 2 + E. Set your sample size equal to 30, and generate,
respectively, X 2 as a normally distributed variable with mean 10 and
variance 25, and E as a normally distributed error term with mean O
and variance 16. The Y values are then obtained from the postulated
regression model. N ext, genera te two more variables, X 3 and X 4 , which,
by construction, bear no relation to Y whatsoever. To do this, generate
X 3 as a normally distributed random variable with mean 5 and variance
16, while X 4 has mean 15 and variance 36. Given these artificial data,
regress (a) Y on X 2 , and (b) Y on X 2 , X 3 and X 4 • You know that,
by design, the latter regression carries a lot of extra baggage due to the
inclusion of two irrelevant variables. Carefully check how the introduc-
tion of irrelevant variables affects the standard errors and coefficient
Model selection and misspecification 219
estimates of the regression coefficient of the relevant variable. What do
you conclude about the effect of including irrelevant variables on the
precision of the estimates of the relevant variable?

Conclusion
This section discussed the implications of omitting relevant variables or
adding irrelevant ones in terms of bias and precision of regression results.
The resulting message is quite complicated. On the one hand, it is impor-
tant that we take seriously the implications of omitting a relevant variable
because it can lead to misleading conclusions as a result of the bias it intro-
duces. But, on the other hand, with collinear variables, adding regressors
to a model has a cost in term of the precision of our estimates due to the
variance inflation factor of the collinear regressor. If a regressor matters a
great deal, the gain obtained by adding it far outweighs the cost of omit-
ting it from the regression. The question of including regressors which are
relevant but less vital in terms of their effect on the dependent variable is
more tricky: the gain in the reduction in bias needs to be balanced against
the loss in precision, particularly when the sample size is relatively small.
The presence of rival proxy variables often amplifies the problem.
Irrelevant variables merely blur the picture, but we do not always know
which variables are irrelevant. The task of variable selection, therefore, is
quite daunting. Let us now see how hypothesis testing can help once we
have decided upon a set of variables to include in our general model.

6.4 TESTING ZERO RESTRICTIONS


It is a poor strategy to start with a regression which only includes a few
variables of direct interest and sets aside other variables which also matter.
In empirical research we cannot assume that other things are equal and,
hence, we cannot set aside important variables merely because they are not
of immediate interest to us. Nor can we discard variables for which there
are strong reasons to include them, although we may doubt whether they
matter a great deal. If you doubt whether one or more of such variables
should be added to a model along with the regressors already included, the
best strategy is to include them right from the start. lndeed, if we begin by
estimating the simpler model which does not include the doubtful variables,
our subsequent testing will not tell us whether or not the model is correctly
specified. The t-statistics, for example, are calculated on the assumption
that the model is correct and, therefore they cannot help us to discover if
in fact it is not. As we saw above, Griffin found a strong and significant neg-
ative statistical relationship between aid and savings. But his results cannot
tell us whether his model is, in fact, correctly specified. To see whether a
simpler model is valid, therefore, it is preferable to start with the general
model and try to see which variables may be dropped through formal
220 Econometrics far developing countries
hypothesis testing. How do we do this?
Suppose, as in the previous section, we have a model with two regressors:
Y; = 131 + l32X2; + ¡33X3; + E; (6.19)
but we doubt whether X 3 affects Y. That is, we have good reasons to
believe the model to be:
(6.20)
obtained by imposing a linear restriction on the general model: namely,
¡3 3 = O. Hence, omitting a variable results from imposing a zero restric-
tion on its coefficient in the general, unrestricted model. Hypothesis
testing then involves that we test the hypothesis H 0 : ¡3 3 = O against the
H 1: ¡3 3 =F O. In this case, you already know how to do this with the t-test.
However, here we shall introduce you to the F-test which is more versa-
tile in its use.

The F-test
The F-test is a powerful tool in specification testing since it enables us to
test a whole range of linear restrictions. The versatility of the F-test, unlike
the t-test, is due to the fact that it does not rely on the standard errors
of individual coefficients, but operates on the residual sums of squares of
the regression as a whole. In other words, the F-test checks whether the
imposition of a linear restriction on a model significantly increases its
residual sums of squares. To do this, the F-statistic takes account of the
degrees of freedom of, respectively, the unrestricted (i.e. without imposing
the restriction) and restricted versions of the model. The general formula
for the F-test is as follows:
F ) = RSSR - RSSu n - ku
(m,n-ku RSS m (6.21)
u
where RSSuR = the residual sums of squares of, respectively, the unre-
stricted and the restricted model estimations; n = sample size; ku = number
of estimated coefficients in the unrestricted model; m = number of linear
restrictions imposed on the model.
The different parts of this formula need sorne explanation. Note that if
we add one or more regressors to an equation, the RSS cannot possibly
increase. If the added variable has any explanatory power at all, it will
reduce the RSS. Consequently, dropping one or more variables from a
model will increase the RSS (unless the estimated slope coefficients of the
deleted variables are exactly equal to zero ). More generally, the imposition
of a linear restriction on the model can never reduce the RSS when the
restricted model is estimated. The numerator of the first term in the formula
shows the difference between these two residual sums of squares, and divid-
ing this by RSSu gives the proportional increase in the RSS from imposing
the restriction. How big does an increase have to be to make it significant?
Model selection and misspecification 221

Box 6.2 The F-distribution


The F-distribution is the distribution of a ratio of two independent
chi-square variables (say, W and Z), each divided by their own
degrees of freedom (say, v1 and v 2 ). In turn, a chi-square distribu-
tion with v degrees of freedom is the sum of squares of v
independent random variables with a standard normal distribution.
Hence:

F = Wlv1
Zlv 2
follows an F-distribution with v1 and v 2 degrees of freedom (of,
respectively, the numerator and the denominator).

Fortunately, it is easy to test whether or not the increase in the RSS


caused by imposing the restrictions is significant, since if we divide the
numerator and denominator of the first term by their respective degrees of
freedom the resulting statistic follows an F-distribution (see Box 6.1) under
the assumption of the null hypothesis. As you know from Chapter 5, the
degrees of freedom for RSSu is (n - ku). Similarly, the degrees of freedom
of RSSR is (n - kR) since it contains fewer coefficients to estimate.
Therefore, the degrees of freedom of the difference term, (RSSR -
RSSu), is given by:
(6.22)
and, hence, the increase in the residual sums of squares as a result of
imposing m zero restrictions on the model will have m degrees of freedom.
More generally, if we impose m linear restrictions on the model, what-
ever their form, the increase in the residual sum of squares will have m
degrees of freedom since the imposition of the restrictions will reduce the
number of coefficients in the model by that number. The resulting F-
statistic is distributed with m, n - ku degrees of freedom.

Exercise 6.3
If the restrictions we impose on a model concern dropping one or more
explanatory variables, both the restricted and unrestricted versions of the
model will feature the same dependent variable. Now, given that R 2 =
ESS!TSS = (1 - RSS!TSS) show how the F-test in equation 6.16 can be
re-expressed using R2s rather than RSSs of the restricted and unrestricted
regressions. To do this, express RSS as a function of TSS and the R 2 , and
substitute this solution in to equation (6.21 ).
222 Econometrics far developing countries
If you completed exercise 6.3, you will have found that the F-test can
also be formulated as:
F = Ru2 - R/ n - ku
(m,n-ku) l _ R 2 m (6.23)
u
Note that in the R 2 version of the F-test, it is the statistic of the un-
restricted model which appears first in the numerator. This is what we
would expect, since Ru2 will be greater than Ri. Note that expression
(6.21) is gene rally applicable, whatever the form of linear restrictions
we impose on the model, while (6.23) can only be used if both the
restricted and unrestricted version of the model feature the same depen-
dent variable.

Dropping a regressor from the model: the F-test and t-test compared
But let us now return to our example in section 6.2 and test formally
whether the variable 1/Y can be dropped from the model. We already have
all the information we require to apply the F-test to Griffin's restriction
imposed on the 'correct' savings model. Griffin imposed the restriction
that ¡33 = O, and, hence, ku = 3 and kR = 2, so that m = l. The sample size,
n, is 66 and we have RSSR = 10,358 and RSSu = 6,637. The restricted RSS
is indeed considerably greater than the unrestricted. But is the difference
significant? We test this with the F-test:
F = 10,358 - 6,637 66 - 3 =
35 3
<1•63 ) 6,637 1 .
This calculated figure compares with a critica! value of just under 4.0
at the 5 per cent level. Since the calculated value is greater than the crit-
ica! value, we can reject the null that ¡3 3 = O, in favour of the alternate
hypothesis that it is non-zero: Griffin's restriction is invalid. We leave it
to you to verify that the R 2 version of the F-test will yield the same result.
You may think that all this is much ado about nothing. If we wish to
test the null hypothesis that ¡3 3 = O, why not just look at the t-statistic?
As can be calculated from the results in equation (6.7), t = 5.92, and,
hence, we reject the null hypothesis at the 5 per cent significance level.
In this case, the t-test is indeed easier, but it is only valid as a test of a
single zero restriction. If we want to impose two or more zero restric-
tions, we can no longer rely on separate t-tests, but we must use an F-test.
In other words, we cannot test the hypothesis that two or more coeffi-
cients in a multiple regression are both zero by looking at their respective
t-statistics; we must do a joint test with the F-test. The reason that
combining individual t-tests is insufficient to test whether we can drop two
or more variables from a model is because, in general, the sampling distri-
butions of these coefficients are not independent of one another: in the
three-variable case, for example, as shown in equation (5.29), the covari-
Model selection and misspecification 223
anee of the least squares slope coefficients is generally not zero, unless
the corresponding regressors are orthogonal. The knowledge that one
coefficient is zero, therefore, will generally affect the probability of the
other being zero as well, and hence a joint test is required.
If our restriction involves dropping only one variable, it does not matter
whether we use the t- or F-tests: both tests are equivalent. In fact, it can
be shown that t2 = F. In our example t2 = 5.922 = 35.1, which, allowing
for rounding errors, is acceptably clase to the calculated F-statistic of 35.3.
But, to repeat, if we wish to restrict more than one variable, we must
apply the F-test.

Testing several zero restrictions at once


A recurrent debate among macroeconomists is whether government
expenditures (particularly government investment) increase or decrease
prívate investment: i.e. crowding in versus crowding out. We explore this
question for the case of Sri Lanka, using the data file SRINA. To start
with, Figure 6.1 gives the scatter plot of prívate investment (/p) against
government investment (/8 ) along with the fitted line which is (t-statistícs
in parentheses):
IP = -648 + l.0418 R 2 = 0.723 (6.24)
(-0.24) (6.84)

32
30
D
28
26 D
D D
24 D D
oo 22 D

e. 20 D
D
E D
(]) 18 D
E
Cii
(])
16
> 14
.!::
(])
(ij 12
>
&: 10
8 D
D D
6
4 D
ºº
2
o
o 4 8 12 16 20 24 28
Public investment ('000)

Figure 6.1 Private versus public investment: Sri Lanka, 1970--89


224 Econometrics for developing countries
At first sight, it appears that, over this period, there was strong crowding-
in: each rupee increase in public investment led to an equivalent rise in
private investment (the coefficient of 1.04 is insignificantly different
from one).
But not so fast. We saw above that if we omit any important variables
from our equation, then our parameter estimates will be biased. Surely
private investment depends on variables other than the level of govern-
ment investment? Our data set includes two further potential determinants
of private investment: the real interest rate (r) and imports (M).
The importance of imported capital to sustaining investment in devel-
oping countries underpins the foreign exchange constraint in the two-gap
model (see Chenery and Strout, 1966): higher levels of imports should
permit higher levels of investment. In Sri Lanka prior to 1977 this
constraint was compounded by quantity restrictions on imports. We use
here total imports, although it may be desirable to exclude consumer
imports from this figure.
The role of the interest rate is more ambiguous. In the standard invest-
ment function a higher real interest rate (cost of capital) results in reduced
investment: this view motivated subsidised interest rates in favoured
sectors in many developing countries for many years. The contrary posi-
tion, associated with McKinnon (1973) and Shaw (1973) (see also Fry,
1988), is that investment is constrained by lack of investible resources,
savings being discouraged by low interest rates: higher interest rates thus
lead to higher investment. Which position is correct is an empirical matter,
so the sign on the real interest rate in our regression is also of interest.
We thus formulate the general model as follows:

Ip,t == f3o + f3i18,t + f32M1 + f33rt + E1 (6.25)


which yields the following estimates:
IP == -12,147 - 0.3318 + 0.61M - 17,970r
(-4.31) (-1.17) (5.26) (-1.54) (6.26)
R 2 == 0.898 RSS == 1.33 x 108

We can see that the coefficient on 18 is now insignificantly different from


zero (at the 5 per cent level): there is neither crowding in or out, a result
which contrasts markedly with our earlier conclusion from the simple
regression.
Which equation should we use to decide on the relationship between
public and private investment? Clearly, equation (6.25) is a general model
of which the simple regression is a restricted version. If these restrictions
are valid we may proceed to test the crowding-in hypothesis on the basis
of the simple regression estimates. But if the restrictions are invalid then
these estimates may not be used. What are these restrictions and how do
we test them? The simple regression omits the variables M and r; that
Model selection and misspecification 225
is, it imposes the restriction: [3 2 = [3 3 = O. We may use an F-test to test
whether or not both these restrictions are jointly valid. First we lay out
the hypothesis being tested, as follows:
Ha: f32 = [33 = O
H 1: at least one of the coefficients, [3 2 or [3 3, is non-zero
The unrestricted equation is given by equation (6.25), so that ku = 4,
while the restricted version is the simple regression with kR = 2. Therefore,
m = 2 (as can be seen from Ha)· Let us use the R 2 version of the F-test.
To do so we make sure to specify the R 2 values up to three decimal places
to avoid excessive rounding errors (which can be quite substantial). The
F-statistic then becomes:
0.898 - 0.723 16
F(2,16) = 1 - 0.898 X 2 = 13.73

which compares with a critica! value of 3.63 at the 5 per cent level.
Therefore, we must reject the null hypothesis that the simple regression
of private investment on public investment is a valid model of the deter-
minants of private investment. We conclude that there appears to be
neither crowding in nor crowding out in the Sri Lankan case. We shall
retum to this example, however, in section 6.7.
At this juncture, we should stress an important point. In both examples
discussed above, Griffin's aid-savings relation and the crowding-in hypoth-
esis, we started with a simpler model and then showed how the results
changed dramatically when additional variables were introduced to the
regression. This arder of presentation was a pedagogical device to illus-
trate the importance of not jumping to conclusions on the basis of
oversimplified regressions. If we had been presenting these regressions as
results from research there would have been no call to report the simple
regression results as they are statistically meaningless (though we may
choose to report that we tested whether the simpler model were a valid
restriction of the data and found it not to be so).

Restricting ali the slope coefficients to zero:


testing the significance of the R 2
We can also use the F-test to test whether all regressors can be dropped
jointly from the regression. If this restriction is acceptable, then none of
the variables matters in explaining the variation in the dependent vari-
able. In this case, the test can best be done using the R 2 variant of the
F-statistic. The reason is that a model without any explanatory variables
apart from the constant term has an ESS equal to O. Consequently,
Ri = O, a result which reduces equation (6.23) to:
226 Econometrics for developing countries
F - Ruz n - ku
(m,n-kul - l _ R 2 m
u (6.27)
There is, however, generally no need to calculate this F-value since most
software packages list its value as the F-statistic of the regression.

Exercise 6.4
With the results of exercise 6.1, formally test whether (a) both explana-
tory variables and (b) the price variable only can be dropped from the
three-variable model. In the latter case, use both the t- and the F-tests
and check the relation between them.

A word of warning: checking model assumptions


Before we engage in variable selection through testing downwards,
however, we should always carefully investigate whether our starting
point, the general model itself, reasonably satisfies the assumptions of
classical normal regression. In other words, we should not take our
general model for granted. The reason is that statistical inferences are
only valid if the model assumptions are reasonably satisfied in practice.
For this reason, we should always subject our general model to diagnostic

40 ...•
20

Q)
o
E
o
(.)
e
~ -20
Cl
e
-~
en
-40

-60

26

-80

o 10 20 30 40 50 60
Aid/income

Figure 6.2 Scatter plot of S/Y against A/Y


Model selection and misspecification 227
Coef = -17377.548, se = 2923.9002, t = -5.94
31.5875 15


5



-66.4916 26

-0.00111 0.002809
e(inv YI X)
s1a1a"'
Figure 6.3 Partial regression: S/Y on 1/Y

misspecification checks and test. There is indeed a real danger that we


may take our general model too easily for granted and jump straight into
hypothesis testing. In fact, as you may have noticed, this is precisely what
we did in the case of Griffin's hypothesis. But we never really checked
whether his general model reasonably satisfies the assumptions of clas-
sical normal linear regression. We make such an analysis by looking at
the various plots (scatter plots, box plots and residual plots).
How <loes Griffin's model fare? In fact, there are quite serious prob-
lems with the model due to the presence of infiuential points and out-
liers. A glance at the scatter plot of S/Y against A/Y in Figure 6.2 reveals
that the simple regression of both variables is not a good fit at all. In
fact, three data points - respectively, points 5, 15 and 26 - dominate the
scene.
But our majar concern is whether the assumptions of the general model
as given by specification (6.6) and estimated in equation (6.7) are reason-
ably satisfied in practice. A simple scatter plot of S/Y on A/Y <loes not
tell us how well the coefficients of the multiple regression are supported
by the data. Figure 6.3 shows the partial regression plot for the slope coef-
ficient of I!Y in the estimated general model. This plot reveals that the
coefficient estimate of l!Y is extremely shaky and depends essentially on
two data points only! These happen to be points 15 and 26 which already
proved troublesome in the simple scatter plot.
228 Econometrics far developing countries
The implication of this discovery, however, is that our earlier F-test (or,
for that matter, t-test) whether the variable 1/Y could be dropped from
the general model rests upon very fragile foundations. Based on this test,
we concluded that the slope coefficient of this variable was significantly
different from zero, but we now see that the result we obtained is solely
based on the position of two data points. In fact, if we remove these two
points from the sample, the t-statistic of l/Y becomes insignificant. The
implication is, therefore, that the test we did earlier to see whether
Griffin's general model could be simplified was largely meaningless. Given
that the assumptions of classical normal linear least squares are clearly
violated in this case, our estimations do not allow us to make meaningful
inferences. The fact, however, that the removal of two data points render
the coefficient of l/Y insignificantly different from zero should not lead
us to conclude that we might as well use the simpler model. A quick
glance at Figure 6.2 shows that the simple regression is not better in terms
of satisfying the assumptions of least squares.
Perhaps this comes as a bit of a surprise. But think how often we accept
hypothesis tests presented in papers, articles or books at face value because
we take it for granted that the assumptions of the model are satisfied. The
problem is that often the authors may have done exactly the same, testing
hypotheses based on the normality assumption without first checking
whether this assumption prevails in practice. This type of hypothesis test-
ing may look good on paper but, if the assumptions are not reasonably valid
in practice, the result of the test is also not valid.

Exercise 6.5
Rework the example of Griffin's hypothesis and systematically check the
assumptions of the model using diagnostic graphs, normality checks,
studentised residuals and DFBETAs.

If you tried out studentised residuals and DFBETAs, you will have seen
that they also confirm the presence of outliers and infiuence in the regres-
sion: the studentised residuals for data points 15 and 26 are, respectively,
2.3 and -6.73, and data point 26 gives a DFBETA value of -2.97 for the
slope coefficient of l/Y. These statistics, therefore, also confirm that
outliers and infiuential points are prevalent and are likely to distort the
inferences we make.
An important lesson we learn from this example is that we should never
forget that the presence of serious outliers and infiuential points is a sign
of misspecification. It tells us that the error term of the model still contains
a lot of meaningful information which we have notas yet grasped. Jumping
ahead into hypothesis testing without carefully checking the assumptions
of the model is a poor strategy in data analysis. Superficially, it may look
good on paper. But many of these inferences may prove worthless if you
Model selection and misspecification 229
care to check your assumptions. Befare drawing conclusions from a model
make sure its foundations are sound so that you can assert them with
sorne confidence.
In this section we have seen how to use an F-test as a means of testing
zero restrictions, so we can test a specific model against a more general
one. But the example we have just seen shows that you must also check
the specification of the general model. If the 'general model' still has
omitted variables then this problem will quite possibly show up in the
residuals - possibly as influential points as in the example just discussed.
But omitted variables can also produce problems of heteroscedasticity
(see Chapter 7) and serial correlation (see Chapter 11). Hence a key
message of this text is borne out here: always look at the residuals for
clues as to possible misspecification.

6.5 TESTING NON-ZERO LINEAR RESTRICTIONS


Imposing zero restrictions on a model implies that we seek to drop one
or more variables from the regression. Non-zero linear restrictions are of
a different nature. Their purpose is not to drop a regressor, but to check
whether one or more coefficients of the model obey sorne linear condi-
tion. Why would we do this? The best way to answer this is to consider
a few examples.
A common example of such a non-zero restriction is to test for constant
returns to scale in a Cobb-Douglas production function:
(6.28)
where Q is output, K capital, L labour and E' a multiplicative error term.
OLS estimation requires linearisation of the equation by taking logs of
both sides:
(6.29)
where ~ 1 is the logarithm of A and E the logarithm of E'.
Now, suppose we want to test the hypothesis that there are indeed
constant returns to scale, i.e. to test the null hypothesis that ~ 2 + ~ 3 = l.
This is a non-zero linear restriction. How do we impose this restriction
on the model? To do this, re-write the null hypothesis as ~ 2 = 1 - ~ 3 and
substitute this expression into the unrestricted model, as follows:
lnQ; = ~1 + (1 - ~ 3 )lnL; + ~ 3 lnK; + E; (6.30)
= ~1 + lnL; + ~ 3 (lnK; - lnL;) + E;
and, after rearranging terms, we get:
ln(Q/LJ = ln~ 1 + ~ 3 ln(K/L;) + E; ( 6.31)
Take a good look at this restricted version of the model. Two things
are noticeable. First, the restricted model now has only two coefficients
230 Econometrics for developing countries

Box 6.3 Constant retoros to scale


Constant returns to scale (CRS) means that an increase in all the
factors of production by a factor m increases output by the same
proportion. Thus if:
Q = ALl32 Kf33
then output with increased factors of production (Q') is given by
Q' = A(mL)i32 (mK)i33

= Qml32+133

Thus Q' = mQ (i.e. there are constant returns to scale) if 13 2 + 13 3 =


1. If Q' < mQ there are decreasing returns to scale and increasing
returns if Q' > mQ.

instead of three in the unrestricted version. The imposition of the linear


restriction, therefore, has reduced the number of model parameters by
one. Second, the dependent variable in the restricted model is different
from that in the unrestricted model. The latter features the logarithm of
output as the dependent variable; the former the logarithm of the output
to labour ratio. Hence, if we test this restriction with real data we cannot
use the R 2 variant of the F-test, but must use the RSS variant instead.

Exercise 6.6
The data file PRODFUN contains data from a developing country busi-
ness survey covering two manufacturing sectors. Estímate specifications
(6.29) and (6.31) using the data in this file, and formally test the hypoth-
esis whether there are constant returns.

Exercise 6.7
Using the data set KRISHNAJI (food price and manufacturing demand
in India), estímate equatíons (5.5) and (5.41), which differ inasmuch as the
latter imposes the restriction of a unitary income elasticity on the demand
for manufacturing goods: that is, 13 2 = 1, which is a non-zero linear restric-
tion. Note that the unrestricted (equation 5.5) and restricted (equation
5.41) versions of the model do not feature the same dependent variable.
1 Use (6.31) to calculate the F-statistic to test the restriction of a unitary
elasticity.
Model selection and misspecification 231
2 Check what would happen if you had calculated the F-statistic with
formula (6.23).
Another common example in economics which involves non-zero restric-
tions is the assumption of no money illusion in demand equations (i.e.
demand depends on real not nominal income). We carne across such a case
in the previous chapter when discussing the demand for manufacturing
goods in India. A further example involves testing for parameter stability
across different samples. This is the issue we shall turn to in the next section.

Exercise 6.8
Khan and Reinhart (1990) argue that the productivities of public and
prívate capital are different, the latter being more productive. To test this
they first regressed growth (y) on aggregate investment (!), growth of
the labour force (L), growth of exports (X) and a constant. They then
repeated the regression with investment desegregated into public (Ig) and
prívate (Ip). Using cross-section data for 24 countries they achieved the
following results:
y= 1.085 + 0.119/ + 0.427 L + 0.212X R2 = 0.660
(0.81) (2.36) (1.33) (4.97)

y= 2.145 + 0.158/P - 0.108/g + 0.573L + 0.163X R2 = 0.737


(1.66) (3.27) (1.02) (1.94) (3.75)
Show how you would use these results to test the hypothesis that the
productivities of public and prívate capital are the same.

6.6 TESTS OF PARAMETER STABILITY


When we run a regression with sample data we assume that the data are
sampled from the same population to which the population regression line
applies. But what if our sample lumps together data which do not really
belong together inasmuch as the same model does not apply to them? It
is often useful, therefore, to test for parameter instability across different
samples before we decide to pool them. Once more, this test can be done
with the help of linear restrictions.

Pooling data
In Chapter 4, we discussed the relation between using the data file
SOCECON. Our model, therefore, is as follows
(6.32)
This model assumes, however, that the values of (3 1 and (3 2 are the same
232 Econometrics far developing countries

80 o
o o o
o o
o 000 00
o o o o
00
o 00
o o o
o o o 00
70 o o o
o
o 0000 o
>. Q() o o o
(J o
e: o
C1l o o o
00 o
tí o o
Q) o
a. 60 <>
X
Q)
o
2 o o
o
:.:J o o
00
o o
o o
o
50 o o o
o o
o 00
o 00 o o
o
o
o
40

4 5 6 7 8 9 10 11
Log (GNP per capita)

Figure 6.4 Scatter plot of life expectancy against log(GNPc)

for all observations. Figure 6.4, which depicts the scatter plot of the data
using all available countries, shows there are actually two sets of rela-
tionships: the scatter is steeper on the left-hand side (left of the vertical
line indicating incomes below $2,500), and fiatter for countries with higher
incomes. This result is not very surprising, since there is a physiological
limit on life expectancy.
The result of these underlying data patterns is that if we fit the regres-
sion line to all observations, it does not capture the relationship at all
well, as is clear from Figure 6.4. The resulting regression line is as follows:
/\.
LE; = 20.96 + 5.80log(Y;) R 2 = 0.75 (6.33)
(2.41) (0.32) RSS = 3125.8
Looking at the scatter suggests that the relationship between income
and life expectancy is different for high and low income countries. That
is, we should have two separate regressions:
For Y; ~ $2,500 LE;= a 1 + a 2 log(YJ + E2 ; (6.34)

For Y; > $2,500 LE; = 'Yi + "{ 2 log(YJ + E3; (6.35)


Running a single regression means that we pool the data: that is, we
assume an identical intercept and slope for the two sets of observations.
Model selection and misspecification 233
More formally this can be stated as the null hypothesis:
Ha: ª1 = 'Y1 ª2 = 'Y2
where the alternate hypothesis entails that at least one of these equali-
ties does not hold.
This null hypothesis may be tested by the application of the usual F-
test. In this case the unrestricted model consists of the two separate
equations (6.34) and (6.35). Hence, the unrestricted residual sum of
squares, RSS u, is the sum of the residual sums of squares from the two
regressions: i.e. RSSu = RSS1 + RSS2, where RSS1 and RSS2 are the residual
sums of squares from the two sub-sample regressions. This version of the
F-test is sometimes called the 'Chow test' (or first Chow test). The R 2
version of the F-test cannot be used for testing parameter stability.
The number of regressors in the unrestricted model, ku, is the total
number of regressors in the two equations: in this case ku = 4. The sub-
sample regressions are:
A
For Y;~ $2,500 LR; = 7.96 + 7.89 log (Y;) R 2 = 0.63 (6.36)
(4.64) (0.72) RSS1 = 2217.6
A
For Y;> $2,500 LE; = 37.3 + 3.94 log (Y;) R 2 = 0.42 (6.37)
(7.25) (0.78) RSS2 = 529.0
Hence, we calculate RSSu = 2,217.6 + 529.0 = 2,746.6. Consequently, the
F-statistic may be calculated as:
- 3125.8 - 2746.6 111 - 4 - 39
F< 2 •106l - 2746.6 ' 2 - 7· (6.38)

This calculated value is much greater than the critica! value of a little over
3.07 at the 5 per cent level, so we reject the null hypothesis that it is valid
to pool our data.
Testing for the validity of pooling is often not done in published
research. Yet a glance at the scatter plot shows in this case that it is
improbable that the data may be pooled for a linear regression. This is
thus yet a further example where preliminary data analysis - such as
looking at the scatter plots - can yield useful information about specifi-
cation ( and warnings about possible misspecification) of your model. What
we can do once the data has rejected pooling is discussed in section 6.7
below. First we show how a similar test may be used to test parameter
stability across time.

Exercise 6.9
White (1992) presents results listed in Table 6.1 for the regression of real
GDP growth on a constant, savings, exports, grants and other capital
inftows for three developing regions: Africa, Asia, and Latin America and
234 Econometrics for developing countries
Table 6.1 Results of growth regressions
Savings Exports Grants Capital infiows RSS
Africa 0.08 -0.02 0.15 0.08 26,401
(0.02) (0.02) (0.03) (0.08)
Asia 0.09 0.03 0.07 -0.17 3,150
(0.03) (0.01) (0.07) (0.08)
LAC 0.18 -0.09 0.08 -0.10 10,290
(0.03) (0.02) (0.04) (0.11)
Ali regions 0.11 -0.03 0.13 0.01 41,717
(0.11) (0.01) (0.02) (0.05)
Note: Absolute values of t-statistics are listed within brackets.

Caribbean. He also gives the results from pooling the data across the three
regions. What assumption is being made in running the pooled regres-
sion? Test this assumption by means of an F-test, stating clearly your null
hypothesis. (The size of the pooled sample is 1,334 observations.)

Parameter stability across time


If we estimate an equation using time-series data we assume that the para-
meters of the model remain constant over time. That is, we are pooling
data from different time periods. Whether or not this is a valid procedure
may be tested in precisely the same way as we tested pooling data above.
A well-known example of a time-series regression is that of the
terms of trade of developing countries against time. Prebisch and Singer
maintained that these were declining over time. There are a number of
theoretical and measurement issues that are important to this debate
which we shall not discuss here (see Sproas, 1980; Sapsford, 1985; Grilli
and Yang, 1988). The data file TOT gives the ratio of an index of the
price of non-fuel primary commodities to that of an index of export unit
values for manufactured goods from industrial countries. A lag-linear
regression using these data gives the following result (standard errors in
parentheses) for the period from 1950 to 1986:
~
ln(TOT); = 0.163 - 0.007t R2 = 0.31 (6.39)
(0.039) (0.002)
The R 2 of 0.31 is not particularly good. A glance at Figure 6.5 tells us
why this poor result is obtained. The data appears to have a break in the
early 1970s. If there are good reasons to believe why a break would occur
at that point, we can refer to it as a structural break. If, however, there
are no good reasons for a break in the trend, the relevant conclusion is
that the pattern we seek to impose on the data is not an adequate fit. In
this case, there are good reasons why the trend would have changed
around this time: the oil price boom in 1973-4 was accompanied by a
Model selection and misspecification 235
1.50

1.40

1.30

1.20

1.10

1.00

0.90

0.80

0.70

0.60
1950 1955 1960 1965 1970 1975 1980 1985

Figure 6.5 Terms of trade of developing countries over time

marked increase in the prices of many other primary commodities. We


really should have looked at the graph of the data before estimating equa-
tion (6.39), since it seems that we need two equations, not one. Let us
split the sample into two sub-samples, 1950-72 and 1973-86.
The model therefore becomes:
ln(TOT) 1 = a 1 + a 2t + E 11 for t = 1950 ... 1972 (6.40)
ln(TOT) 1 = -y 1 + -y 2t + E21 for t = 1973 ... 1986 (6.41)

The single regression we estimated above (equation (6.39)) imposes the


restriction on the data that:
Ha: ª1 = 'Y1 ª2 = 'Y2
This restriction is exactly the same as the one we encountered in testing
the validity of pooling the data. And so, as before, we test this null ( against
the alternative that at least one of the equalities does not hold) with the
F-test where the unrestricted residual sum of squares is the sum of the
RSSs from the two sub-sample regressions. Estimating equations (6.40)
and (6.41) gives the two residual sums of squares as:
RSS1 = 0.083 RSS2 = 0.121
These results may be used to derive a calculated F-statistic of 22.0, which
236 Econometrics far developing countries
is greater than the critical value of F(2,33 ) (just under 3.32) at the 5 per
cent level, so we reject the null hypothesis that it is valid to pool the data.

Chow's second test and predictive failure


lt may happen that the supposed structural break occurs late in the time
series. Especially where we have a large number of regressors it is possible
that we have too few observations to estimate the equation for the second
sub-sample. In this case, Chow suggested an alternative test (also based
on the F-statistic) which is:
RSSR - RSS1 n - kR - m
F(m,n-kn-m) = RSS (6.42)
1 m
where RSSR is, as before, the RSS from the estimated equation for the
whole sample period, RSS1 is the RSS from the estimated equation for
the first sub-sample, kR is the number of regressors (including the constant)
in the restricted equation, and n is the number of observations in the full
sample. That is all as before. However, m is no longer the number of
restrictions, but is now the number of omitted observations (i.e. the
number of observations is the smaller of the two sub-samples). The null
hypothesis is as before.
Chow's second test can be used as a predictive failure test and is a very
useful diagnostic tool in testing for misspecification. In fact, in small
samples it is often not so difficult to find a reasonably good fit. More
demanding than getting a high R 2 is to be able to predict out-of-sample
values. If the model really is the data generation process, then it should
be able to do this (subject to error ftuctuations). A useful test, therefore,
is to withhold observations (one or two will suffice, though up to four is
preferred if possible) from the data set when estimating and then compare
actual with predicted values. The relevant test statistic is the F-statistic
calculated by the formula in equation (6.42).
Whilst the discussion of Chow's second test has been illustrated in the
context of time-series data, there is no reason why it cannot be applied
to cross-section data. In the latter case, make sure that the observations
to be omitted appear at the end of the data file.

Exercise 6.10
Repeat the test for a structural break in the developing country terms of
trade putting the break at 1974, rather than at 1973 as is done in the text.
Which is the more appropriate break point?

Exercise 6.11
Population figures for Kenya for the period 1968-88 were regressed on
a constant and a trend. The residual sum of squares (RSS) from this
Model selection and misspecification 237
regression was 8.77. The same equation was re-estimated for the sub-
sample 1968-85, from which the RSS was 3.89. Use a Chow test to
determine the stability of the parameters.

6.7 THE USE OF DUMMY VARIABLES


Take another look at our regression of the terms of trade against time.
We found that they declined at 0.7 per cent per annum over the full sample
period, yet we established that the parameters are not stable. As we can
see from Figure 6.5, the effect of the jump in 1973-74 is to 'level out' the
fitted line - the slope coefficients from the two sub-samples suggested a
much more rapid decline. Can we model this structural break in the early
1970s in a single equation? Yes, we can. Dummy variables allow us to
vary either the intercept or slope coefficients between sub-samples. Let
us first illustrate the use of intercept dummies.

Intercept dummies
An intercept dummy is a variable which takes the value one for a specific
sub-sample and zero for the rest of the sample. In our example on the
decline in the terms of trade, we shall distinguish the periods 1950-72, for
which years the variable D UM takes the value zero, and the remaining
years (1973-86) for which it takes the value l. The regression equation
beco mes:
(6.43)
Note that DUM must also be given a time subscript since it is a variable.
Now, when DUM = O, the estimated regression line is given by:
~
ln(TOT) 1 = ¡3 1 + ¡3 3t (6.44)
But when D UM = 1, the line is:
~
ln(TOT) 1 = (b 1 + bz} + b3t (6.45)
The coefficient b2 is the differential intercept - that is, the shift in the
intercept for those observations for which D UM = l. Do not interpret b 2
as the intercept for the second sub-sample - this is not what it is. The
intercept for the second sub-sample is given by b1 + b2•
Estimating equation (6.43) using the terms of trade data set yields:
~
ln(TOT) 1 = 0.254 + 0.262DUM - 0.017t R2 = 0.57 (6.46)
(0.038) (0.059) (0.003) RSS = 0.299

The slope coefficient now shows an annual deterioration in the terms of


trade of 1.7 per cent - two anda half times more than befare! Excluding the
intercept dummy resulted in omitted variable bias, giving an underestimate
238 Econometrics for developing countries
1.50 ~---------------------------...,

1.40

1.30

1.20

1.1 o

1.00

0.90

0.80

0.70

Figure 6.6 Fitted line of terms of trade with intercept dummy

of the true slope coefficient. Figure 6.6 shows the fitted lines from equation
(6.46). You will have observed that we have here assumed that the slope is
constant between sub-samples, but that the intercept can vary. So we are still
imposing a restriction on the unrestricted (two-equation) model. It is left to
you as an exercise to test this restriction. You will find that it is not valid.

The intercept dummy interpretation of a studentised residual


What happens if we construct a dummy such that it selects one only obser-
vation in the sample? Hence, the dummy will equal 1 for this data point
only and zero otherwise. This is a very special case of an intercept dummy:
we divide the data into two sub-samples, one of which has only one obser-
vation in it. Why do we do this?
If we include this type of one-point dummy in a regression, the residual
for this data point will be zero. The reason is simple: the estimated coeffi-
cient of the dummy variable will make sure that the point 'fits' the regres-
sion. Consequently, this data point will not play any role in determining
the other coefficients of the model. Hence, using a one-point dummy is
equivalent to deleting the point from the regression and estimating the
coefficients of the model with the remaining data points. The slope coeffi-
cient of the dummy is nothing else but the vertical deviation of its data point
from the regression line estimated with the other points in the sample. The
Model selection and misspecification 239

0.25 75

0.2

Q) 0.15
.a
'6
e
Q)
0.1
c.
X
Q)
0.05
E
~
:::; o
<..l
~
.r::
j -0.05
e
" -0.1

-0.15

-0.2
-0.2 -0.15 -0.1 -0.05 o 0.05 0.1 0.15 0.2 0.25
Growth: recurren! revenue
srara•M
Figure 6. 7 Regressions of RE on RR with and without data point 1980

Table 6.2 Regressions with and without intercept dummy for 1980 (t-statistics
in brackets)
RE on RR D80 RSS R2
1 Ali observations 0.934 0.0667 0.788
(8.405)
2 Ali except1980 1.047 0.0347 0.890
(12.05)
3 All observations 1.047 0.189 0.0347 0.890
(12.05) (4.077)

estimated coefficient of the dummy, therefore, is a deletion residual.


The studentised residual then is nothing else but the t-statistic of the
coefficient of the dummy variable. lt tests whether the distance of the
data point from the regression line estimated without including it is signif-
icantly different from zero. If so, it would indicate that this data point
might be an outlier. This explains why, in Chapter 4, we hinted that studen-
tised residuals were in fact t-statistics.
To illustrate the point, consider once more our example in Chapter 4 on
the relation between the growth rates of recurrent expenditures (RE) and
revenues (RR) in Tanzania, 1970-90. There we noted that the year 1980 was
quite exceptional because it signalled the onset of a majar economic crisis.
In this year, the Tanzanian government appears to have been caught
240 Econometrics for developing countries
unaware when trying to maintain government expenditures while its rev-
enues collapsed as a result of the crisis. Consequently, as we saw in Chapter
4, the data point for 1980 (corresponding to the growth rates from 1979 to
1980) was both an outlier and an infiuential point. Let us see how we can
obtain the standardised residual for this data point using an intercept
dummy, D80. Hence, D80 = 1 in 1980, and zero otherwise. Table 6.2 gives
us three regressions, all without constant term: (a) RE on RR with all data
points; (b) RE on RR with all data points except 1980; and (c) RE on RR
and D80. Figure 6.7 shows the scatter plot along with the regression lines
corresponding to regressions 1 and 2 in Table 6.2.
Take a careful look at this table. First, the second and third regression
have identical values for, respectively, the slope coefficients of RR, its t-
values, the RSSs (residual sums of squares), and the R 2 s. This shows that
including an intercept dummy which selects one observation from the
sample is equivalent to running the regression which does not include the
data point. Hence, the slope coefficient of the dummy in the third regres-
sion ensures that the residual for 1980 equals zero. Consequently, the
slope of the dummy measures the vertical distance of the data point for
1980 with respect to the regression line estimated without including 1980
in the sample (regression 2). The t-statistic for this slope coefficient is the
studentised residual of the simple regression of RE on RR. If this slope
turns out to be significantly different from zero, it indicates that the corre-
sponding data point (1980) is an outlier. In this case, the t-value of 4.077
is well above the critica! limit and, hence, leads us to reject the hypoth-
esis that the population value of the dummy's slope equals zero. The test
confirms that 1980 was an exceptional year inasmuch as it signalled the
onset of Tanzania's economic crisis of the early 1980s.
Note that the data point for 1980 is not just an outlier but also exerts
sorne infiuence on the regression line. This explains why the slope coef-
ficient in regression 1 is less than that in the other two regressions. As
you can see from the graph, the point 1980 pulls the line slightly in its
direction. Table 6.2 also shows that the RSS of regression 1 is nearly
double the RSS of the other regressions which shows that the outlier has
quite a heavy weight in the residual sum of squares.
In this example we expected beforehand that the data point for 1980
might give us problems and, hence, we formally tested this by including
the intercept dummy for that particular year. But in many applications
we do not have any a priori reason to believe that any point is an outlier.
Generally, we compute all studentised residuals and see if any of them is
exceptional. In this case, however, we cannot just use the normal 5 per
cent significance level to pick the critica! value of the t-statistic. Why not?
The reason is quite simple, but not immediately obvious. Take an
example. If the probability of success in any given trial of a game is 5 per
cent, the probability of encountering one success in ten successive (inde-
pendent) trials will not be 5 per cent. Similarly, if we seek to test whether
Model selection and misspecification 241
one and only one data point is an outlier, the 5 per cent significance level
is appropriate far the corresponding t-test. But if we look at a string of
n residuals at once to test whether there is an outlier among them, we
cannot use the critical t-value corresponding to a significance level of 5
per cent on a single trial. Indeed, in this case, n different t-tests are essen-
tially being made. These multiple tests cannot be treated farmally as if
they were separate tests, but instead, in arder to end up with a signifi-
cance level of no larger than 5 per cent far the test that one of these
residuals is an outlier, we need to use the t-test with a significance level
of 5 per cent divided by n, the sample size (Myers, 1990: 225). In our
example, the relevant critical t-value is 4.29 far 20 observations with two
coefficients in the model. Given the small probabilities involved, most
tabulations of t-tables do not give us critical values to do this type of
outlier test (see, however, Myers, 1990: Appendix C). However, you
should not worry if you do not have access to a detailed t-table. In Chapter
4 we argued that studentised residuals can best be used as exploratory
diagnostic tools. The point we want to make here is that you should not
use the normal 5 per cent critical t-values to pick out seemingly large
studentised residuals since there is nothing exceptional about having a
relatively large value now and then in a bigger sample.

Slope dummies
An intercept dummy allows us to vary the intercept across sub-samples
while keeping the slopes of the regressors constant. But at times we want
to be able to vary the slopes. This requires the use of slope dummies. To
see how we can construct these, let us continue far a while with this
example of the simple regression of Tanzania recurrent expenditures
against recurrent revenues. Recall that, in Chapter 4, the exploratory band
regression of RE on RR noted a mild non-linearity of the regression curve.
More particularly, with the exception of the outlier 1980, it appeared as
if the slope coefficient was lower when the growth rates in recurrent
revenues were positive than when they were negative. Let us see how we
can use a slope dummy to check whether this is true.
To obtain a slope dummy, we first construct an intercept dummy,
DNEG, such that DNEG = 1 if RR < O , and O otherwise. The dummy
DNEG, therefare, picks out all observations far which the growth in recur-
rent revenue was negative. The slope dummy, DRR, is then obtained as
fallows: DRR = DNEG.RR. Hence, DRR = RR when RR < O; otherwise,
DRR = O. We now farmulate our regression modelas fallows:
(6.47)
As befare, the regression model features no constant term. Now, if
DRR =O, the slope of the regression will be 13 2 , but if DRR = RR (when
RR < O), the slope coefficient will be (13 2 + 13 3).
242 Econometrics for developing countries

0.25

0.2
l!:? 0.15
.a
'5
e:
Q)
a.
0.1
X
Q)

E 0.05
l!:?
;:;
(.) o
l!:?
.e -0.05
~
e
(9 -0.1

-0.15

-0.2

-0.2 -0.15 -0.1 -0.05 o 0.05 0.1 0.15 0.2 0.25


Growth: recurren! revenue
STi3Ti3"'

Figure 6.8 Scatter plot of expenditure against revenue

Estimating equation (6.48) yields the following results (t-statistics in


brackets):
RE1 = 0.994RR 1 + 0.2l7DRR1 + 0.2l7D801 ; R 2 = 0.90 (6.48)
(9.99) (1.08) (4.1) RSS = 0.0325
which yield two regressions, both through the origin (i.e. they have the
same intercept), as shown in Figure 6.8. As you can see, the slope of the
line when RR is positive is 0.993 (about 1), while it is (0.993 + 0.217), or
1.21, when RR is negative. Hence, at first sight, there is a marked differ-
ence between both slopes. However, in Chapter 4 we suggested that this
kink in the regression line did not appear to invalidate the linearity
assumption for the data as a whole (with the exception of the outlier year
1980). Does our regression confirm this view?
The advantage of using dummies is that we can use the familiar t-test
to see whether the same regression applies to all data points or not. In
this case, the t-value of the slope dummy is 1.08, a value which does not
lead us to accept the alternative hypothesis that the population coefficient
of slope dummy is significantly different from O. Hence, there appears to
be no reason to split the sample for negative and positive values of the
growth rate of government revenue. The overall regression (excluding the
outlier year 1980) through the origin appears to be a good enough
summary of the data.
Model selection and misspecification 243
32
30 D
28
26 D
D D
24 D D
oo 22 D

E 20 D
D
E D
Q) 18 D
E
¡;; 16
Q)
> 14
.!:
Q)
ia 12
> ¡JJ
et 10
8
D D
6 D D
4 D

2
o 10 12 14 22 24
6 8 16 18 20 26
Public investment ('000)
Figure 6.9 Sri Lankan investment function with intercept dummy only

32
30 D
28
26 D
D D
24 D D
oo 22
D

E 20 D
D
E D
Q) 18 D
E
¡;; 16
Q)
> 14
.!:
Q)
ia 12
>
et 10
8
D D
6
4 D
ºº
2
o
o 4 8 12 16 20 24 26
Public investment ('000)

Figure 6.10 Sri Lankan investment function with slope dummy only
244 Econometrics far developing countries
Combining intercept and slope dummies: the Sri Lankan investment
function
Let us now return to the Sri Lankan investment function and see how
intercept and slope dummies can be combined to throw more light on the
crowding-in or crowding-out issue. Take another look at Figure 6.3:
government investment appears to fall into two distinct sub-samples - one
set to the right in excess of Rs 16,000 million and another of lower value
in the bottom-left of the graph. Looking at the data set ( or adding data-
labels to our graph if it has this facility) we can see that the former, higher
values all correspond to the years 1979-89 and the latter to 1970-8. The
Sri Lankan government embarked on a liberalisation programme late in
1977 and on an investment boom in 1978-9. This suggests that we could
use dummy variables to distinguish between periods with different policy
regimes. The question then is whether there was a structural break in our
investment function between these two periods? The scatter plot suggests
there was such a break: with crowding in during the earlier period, but
crowding out later on - so the coefficient on government investment
should be positive in the first period and negative later on.
For pedagogical reasons, let us compare the effects of three different
uses of dummies: Figures 6.9, 6.10 and 6.11 show the fitted lines from
introducing, respectively, an intercept dummy, a slope dummy and both
slope and intercept dummies into the simple regression. If we use an ínter-

32
30 D
28
26 D
24
D
oo 22
2- 20 D
D
E D
CD 18 D
E
(¡) 16
CD
.so> 14
.El 12
Cll
>
10
et D
8
O D
6 D
4 D

2
o 1o 12 14 16 18 20 22 24 26
6 8
Public investment ('000)

Figure 6.11 Sri Lankan investment function with intercept and slope dummies
Model selection and misspecification 245
cept dummy only, there is neither crowding in or out in either period
(Figure 6.9). The problem is that, in this figure, we are constraining the
slope coefficient to be the same in each period. In the case in which we
allow the slope to vary (Figure 6.10), we find crowding in to be stronger
in the second period than the first. Here we are imposing the constraint
that the intercept must be the same for the two sub-samples. The result
is that both regression lines look awkward.
As shown in Figure 6.11, it is only when we allow both slope and inter-
cept to vary that we get a clearer picture: the fitted lines now conform to
the pattems apparent in the scatter. This example allows us to draw an
important conclusion: the difference in the slope coefficient between the
two sub-samples will only reveal itself if we also introduce an intercept
dummy. The intercept dummy is clearly necessary since it is not possible
to have a negatively sloped line through the 1979-89 sub-sample of data
points that has the same intercept as a line with a positive slope through
the 1970-8 data points. The intercept dummy, therefore, is necessary to
accommodate the difference in slope. This is, in fact, nothing more than
a specific example of omitted variable bias.
When we include both slope and intercept dummies then both coeffi-
cients vary between sample periods. That is, including a dummy for all
regressors yields the same coefficients as estimating separate sub-sample
regressions. We may therefore test the validity of pooling data either by
the Chow test presented earlier or by an F-test of the restriction that the
coefficients on all dummy variables are jointly zero. These two tests are
equivalent, i.e. they yield exactly the same result. You are asked to verify
this equivalence in exercise 6.13.
This should not lead us, however, to interpret the intercept dummy as
reflective of a higher level of 'autonomous' prívate investment in the later
period. In general, in the presence of a slope dummy, an intercept dummy
should be interpreted as accommodating the change in slope.

Exercise 6.12
Using the example of the regression of life expectancy on the logarithm
of GNP per capita, re-estímate the model using intercept and slope
dummies combined. Compare your results with those obtained earlier with
two separate regressions. Test formally whether either one or both of the
dummies can be dropped from the model. Do your results confirm the
conclusions reached earlier?

Exercise 6.13
Test whether it is valid to pool the data for the regression of IP on Ig using
the data in the file SRINA by (a) running separate sub-sample regres-
sions for 1970-8 and 1979-89; and (b) using dummy variables. Compare
246 Econometrics far developing countries
the estimated coefficients and restricted and unrestricted RSSs from the
two methods, and comment on your results.

6.8 SUMMARY OF MAIN POINTS


1 Omitting relevant variables from a model will bias the estimates of all
the coefficients in the equation. Its impact on the standard errors of
the regression coefficients is more ambiguous: the theoretical standard
errors will be less in the simpler model if the omitted variables are
collinear with the regressors included in the model, but the estimated
standard errors will generally be less in the larger model if the omitted
variables have a sizeable impact on the dependent variable and if the
sample size is not very small. Hence, at times a trade-off exists between
bias and precision when dealing with omitted variables.
2 The inclusion of irrelevant variables does not result in bias, but affects
the precision of all the model's estimates adversely.
3 Conventional hypothesis testing (t- or F-tests) is based on the assump-
tion of correct specification and, hence, does not help us discover
whether important variables have been left out of the picture. The
results may look good, therefore, but coefficient estimates may be seri-
ously biased. This fact provides the rationale for testing downwards,
i.e. general to specific modelling. If you seek to prove that a simpler
model is an adequate representation of the data, start by estimating
the more general model and test down to see whether your assertion
is justified. Merely estimating the simple model will not prove your
point, even if the results look good.
4 Before testing downwards, always check whether the general model
satisfies the assumptions of classical linear normal regression. If not,
your inferences drawn from testing downwards may not be valid.
Always remember that statistical inferences are as sound as the foun-
dations upon which they rest.
5 In combining numerical summaries and diagnostic plots (particularly
partial regression plots), use residual analysis and influence diagnos-
tics to subject your general model to serious scrutiny.
6 Linear restrictions on a general model can be tested with an F-test.
Common restrictions are (a) excluding variables (zero restrictions); (b)
non-zero restrictions (e.g. equality of two parameters); and (c) para-
meter stability.
7 The use of the F-test to test for parameter stability is often called the
Chow test. Chow also provided an alternative specification of the F-
test to apply in the case when one sub-sample has too few observations
to be estimated.
8 Dummy variables, both intercept and slope dummies, may also be used
to model parameter instability. If a dummy is included for every coef-
ficient, then the F-test on the null hypothesis that the coefficients on
Model selection and misspecification 247
the dummies are jointly zero is equivalent to the Chow test. Dummies
allow us to estimate a single equation which differentiates regression
results with respect to different sub-samples. Subsequent testing allows
us to check whether parameter instability across sub-samples applies
to all coefficients alike.
9 Studentised residuals are nothing but t-statistics of dummy variables
which, each in turn, pick out a single observation from the sample.
This is a special application of an intercept dummy.

ADDITIONAL EXERCISES

Exercise 6.14
Using the data in data files INDONA and SRINA test whether it is valid
to estimate a pooled consumption function for the two countries and
comment on your findings.

Exercise 6.15
In section 6.7 we found a structural break in the simple private invest-
ment function for Sri Lanka. But in section 6.4 we found the simple
regression to be misspecified owing to omitted variables. Using the data
in the file SRINA:
1 Construct partial scatter plots of IP against each of lg, M and r.
2 Use your scatter plots to judge if there is a structural break in any of
these relationships.
3 Hence, define a general investment function which regresses IP on lg,
M and r and any appropriate dummy variables.
4 Try to obtain a more specific model but testing restriction on the coef-
ficients of the general model.
How do you interpret your findings?

Exercise 6.16
Mosley et al. (1991) are concerned to examine the impact of adjustment
policies on a range of macroeconomic performance variables (such as
growth of GDP and exports). They estimate a number of equations in
which the main regressors are current and lagged adjustment-related finan-
cia! flows and measures of compliance with conditionality. A small number
of other variables are included (all unlagged). The authors argue that:
The equations represent somewhat crude hypotheses regarding the
determinants of the five dependent variables. There are many other
independent variables which could have been included as explanatory
248 Econometrics for developing countries
variables in the equations. In addition, lags could have been introduced
to more of the independent variables ... However, since it is specifically
the impact of Bank finance and policy conditions which we wish to quan-
tify, we have refrained from more complex specification of the equations.
(Mosley et al. 1991: 210)
Comment on their argument.

Exercise 6.17
Given the following estimates of the consumption function, both calcu-
lated from the same data set of 20 cross-country observations, what would
you expect to be the sign of the correlation coefficient between income
and the real interest rate:
e= 0.14 + 0.82 y - 0.04r

e = o.os + o.83 Y
where C is consumption, Y income and r the real interest rate? Explain
your answer.

Exercise 6.18
Suppose that the true model of prívate investment (Ip) is:
Ip,t = !31 + !3zlg,t + ¡33M1 + Et
but that you estímate
Ip,t = !31 + !32Ig,t + E't
What would you expect to be the direction of the bias in the coefficient
on Ig in the simple regression of IP on Ig? Use the data for Sri Lanka
(SRINA) to verify your answer.

Exercise 6.19
Using the data in data file PRODFUN estímate a separate production
function for each sector. Is it valid to pool the data from the two sub-
sectors? How do your results affect your answer to exercise 6.6?
Part 111
Analysing cross-section data
This page intentionally left blank
7 Dealing with heteroscedasticity

7.1 INTRODUCTION
Real data do not conform to the idealised conditions of the classical linear
regression model. As we pointed out in Chapter 4, you are likely to
encounter heteroscedasticity frequently in economic data, particularly with
cross-section data. The reason is that the variation in the dependent vari-
able seldom remains constant when the level of one (or more) explanatory
variable(s) increases or decreases. For example, not only is the level of
consumption of the rich much higher than that of the poor, but it is also
more varied. The poor have few options but to spend their income on
the basic essentials of life; the rich enjoy the privilege of making choices.
Similarly, there tends to be much less variation in output or expenditure
levels among small enterprises than among large firms. The implication
for statistical analysis is that you will not be able to apply the regression
model to the data straight away. But fortunately, the techniques of trans-
formation make the application of the model possible in very many
situations in practice. In Chapter 4, we showed how a well-chosen trans-
formation can help to convert a non-linear relationship into a linear one.
Like non-linearity, heteroscedasticity is also often due to the skewness in
the distribution of the variables under study. As a result, a suitable trans-
formation can make the heteroscedasticity disappear while making the
average relationship linear at the same time. However, you may not always
be able to do this. There are also cases where the relationship will look
clearly linear but the scatter plot indicates heteroscedastic errors.
If all the other assumptions of the regression model, i.e. the assumptions
of linearity of the regression, independence and zero expectation of the
error term, are valid, then it can be shown that heteroscedastic errors do
not affect the unbiasedness of the least squares estimates of the regression
coefficients. But the precision in estimation of the coefficients is no more
the best. In other words, the least squares estimators are not best linear
unbiased estimators (BLUE) but only linear unbiased estimators. Further,
the standard formulae for the standard errors will not be valid, since they
are based on the assumption of homoscedasticity. Consequently, it is not
252 Econometrics for developing countries
possible to perform the t-tests and F-tests under heteroscedasticity. Thus,
in order to make reliable inferences from the heteroscedastic data on the
basis of the linear regression model, we seek to eliminate heteroscedasti-
city by means of a suitable transformation. Incorrect functional form is the
type of model misspecification most likely to account for heteroscedasticity
in the residuals; but it may also be a symptom of omitted variables.
There is a very important point here, one on which we depart from
many traditional textbooks. Heteroscedasticity (and autocorrelation, dealt
with in Chapter 11) is a violation of our assumptions about the error term,
which has adverse implications for least squares estimation. But we do
not know the errors, but proxy them with the residuals. The residuals are
a function of our model specifications. Hence 'problems' which appear in
the residuals, such as heteroscedasticity, are just as likely to be a result
of a misspecified model as they are of a genuinely heteroscedastic error
in the true model. When coming across a problem in the residuals the
first course of action must always be to check the model specification.
Only once you are sure that the model is correctly specified should you
turn to one of the traditional 'cures' for residual heteroscedasticity.
In Chapter 4, we showed how the residual versus predicted plot can be
used to check for heteroscedasticity. In this chapter, in section 7.2 we
discuss two other plots which are useful in the visual examination for
heteroscedasticity. Next, in section 7.3, we introduce a selection of statis-
tical tests for heteroscedasticity. Section 7.4 discusses how to explore
suitable transformations for elimination of heteroscedasticity in order to
obtain the best estimators. Section 7.5 shows how weighted least squares
regression can sometimes be used to make inferences when the error term
is heteroscedastic. We show that weighted least squares can also be done
by a linear regression with suitably transformed variables. Finally, section
7.6 gives a summary of the main points.

7.2 DIAGNOSTIC PLOTS: LOOKING FOR


HETEROSCEDASTICITY
Under heteroscedastic conditions, the simple linear regression model of
Y on X is given by:
Y = ~1 + ~2 X + Ex (7.1)
where V(Ex) = a}, the conditional variance of Ex given X, varies with X,
with all other assumptions of the classical model remaining the same. The
best way to spot the presence of heteroscedasticity is to look for it using
visual displays. The basic principles to construct such displays are:
1 Suppose a} increases with X.
2 Then, for i > j we will have a} > af.
3 Nov.r, the variance a} is the average squared deviations of the errors
from its own mean at a given value of X.
Dealing with heteroscedasticity 253
4 Consequently, a} > a/ implies that, on average, the deviations of
errors from its own mean will be larger at X= X; than at X= X¡. But
since mean of the errors Ex is zero at every given X, the errors them-
selves are the deviations.
5 Therefore, a} > a/ implies that the magnitude of the errors E;S will
be larger than Eis.
6 Thus, if a} increases with X then the magnitude of the errors will
tend to increase with X.
7 However, the errors Ex are unobservable, but the residuals ex from the
fitted regression serve as estimates of the errors Ex.
8 We can, therefore, expect that the plot of the residuals against the
corresponding values of X will indicate an increasing spread of the
residuals if a/ increases with X.
9 Further, since the fitted Y values are linear functions of X, the same
pattern of increasing spread of the residuals should be discernible in
the plot of the residuals against the predicted values of Y from the
fitted regression.
The above is the basis of the residual versus predicted value plot. But
since the magnitude of the residuals increases with predicted Y, it follows
that the squared and the absolute residuals will also increase with
predicted Y. In general, if a} varies with X then the scatter plots of the
raw, squared and absolute residuals against the predicted values of Y will
all reveal a pattern of heteroscedasticity. At times, however, one of them
is more revealing than the other. It is useful, therefore, to examine all
three plots separately. Let us illustrate their utility.
In the case of the simple regression, the plots of residuals (raw, squared
or absolute) against the X values and the predicted Y values are equiva-
lent. In the case of the multiple regression, however, the situation is more
complex. As a rule, you should first look at the plot of the residuals against
the predicted values for any indication of heteroscedasticity. If there is
any such indication then you should try the plot of residuals against each
of the explanatory variables to find out which of the regressors accounts
for the heteroscedasticity. Generally, there is only one such explanatory
variable. However, in rare situations, heteroscedasticity may be related to
more than one explanatory variable, which makes the life of an empir-
ical analyst more difficult.

An example: urban weekly earnings against the age of workers


The file INDIA contains data on weekly earnings of 261 workers along
with their age, obtained from a survey of workers' households in an indus-
trial town in southern India. The estimated regression of weekly wage
income on age is given below (standard error in brackets),
Income = 8.65 + 4.88 Age R2 = 0.20 (7.2)
(21.13) (0.61)
254 Econometrics far developing countries
Figure 7.1 presents the residual versus predicted plot. The points spread
vertically wider and wider with the increase in the predicted value of
weekly wage income, indicating heteroscedasticity.
Figures 7.2 and 7.3 are the plots of squared and absolute residuals
against the predicted values, respectively. Both these plots exhibit similar
patterns to those observed in Figure 7.1. However, in this example, the
increasing spread of the points is much more clearly discernible in Figures
7.1 and 7.3 than in Figure 7.2. This is because of the presence of one
outlier which is easily identifiable in all three plots. This outlier makes
the magnitude of its squared residual so large compared to the rest of the
residuals that the latter points are pressed down to the horizontal axis in
Figure 7.2. It is always use ful to look at all three plots, as it is possible
that one of them many fail to bring out the pattern of heteroscedasticity,
if any, in a given situation. In this case, plot 7.2 mainly draws our atten-
tion to the presence of an outlier (which, incidentally, is a good way to
spot outliers).
The plots of absolute and squared residuals against predicted values of
Y are not just useful to spot the presence of heteroscedasticity, but, as
we shall see, they also provide valuable clues to the possible form of trans-
formations which may enable us to cope with heteroscedastic situations.
Before we do this, we discuss in the next section how these visual tech-
niques can be complemented by formal testing for heteroscedasticity.

o
Q) o
E o o
o
(,J
.!::
Q) o
Ol o
t1l
:;: g 8 ººo o
ro:::i o
ººo o
o ºo
o o
o o 8@ o
-o o o ºº il o ~ o ¡¡ o
·¡¡:; o oºo o o ºo
Q)
a: 8
0 go o0 o g 0
o oºº
000808§ oºº
o o o o
o o
8 8
o
o 8
o
8 o
o o
o

52.3633 374.897
Predicted wage income

Figure 7.1 Residual versus predicted plot


Dealing with heteroscedasticity 255

506711 o

(/l
Cii
::J
"O
·¡¡;
~
"O
~
Ctl
::J o
O"
(/) o
o o

o
o o
o§ 0
00 o o o
ºº 8 º ° º
o ~ 2 ~ºI 8° ~ 88 ªº o ~
soo
oo@@oo@iiooOo§O~eg89oifio ooioMº o0 B oo ~º
0
0.728144 io8° 0 @ o

52.3633 374.897
Predicted wage income
srara™
Figure 7.2 Squared residuals versus predicted

711.837 o

o
o
o o

o
o o
o o o
8
o o
o@ o@
ooo oo
8 §
o 8ºº o
o o
o o
0.853314
62.3633 374.897
Predicted wage income
srara™
Figure 7.3 Absolute residuals versus predicted
256 Econometrics for developing countries
Exercise 7.1
In Chapter 4, while discussing transformations towards linearity, we used
three examples with data taken from the data file SOCECON: respec-
tively, the relation between energy consumption (E) and GNP (Y), both
per capita; between energy consumption (E) and the degree of urbanisa-
tion as measured by the percentage of the population living in urban areas
(U); and, finally, between life expectancy (L) and GNP per capita (Y).
For each of these simple regressions between the raw data, compare the
plots of raw, absolute and squared residuals against the predicted values
of the dependent variable or against the regressor. In each case, check
which plot is most revealing in terms of detecting heteroscedasticity.

Exercise 7.2
Use the INDIA data set to estímate the regression line between the loga-
rithm of wage income and the age of the worker, compute the residuals,
and plot the raw, absolute and squared residuals against the predicted
values of wage income and against the age of workers. What do you
conclude about the presence or absence of heteroscedasticity?

7.3 TESTING FOR HETEROSCEDASTICITY


There are a number of tests available in the literature which help us to
detect the presence of heteroscedasticity. We shall not review these tests
exhaustively. What follows is an illustration of a selection of commonly
used tests. However, none is foolproof in detecting heteroscedasticity.
Detection of heteroscedasticity really requires a blend of visual and confir-
matory (i.e. hypothesis testing) analysis.
There are two distinct approaches to testing for heteroscedasticity. One
approach is to try to detect heteroscedasticity by comparing the condi-
tional variances across different ranges of the explanatory variable(s).
Bartlett's test and the Goldfeld-Quandt test illustrated below are exam-
ples of this approach. While Bartlett's test is a general test for equality
of variances between groups which can be adopted for the purpose of
testing for heteroscedasticity, the Goldfeld-Quandt test is specifically
developed for this purpose. The other approach is based on the formali-
sation of the graphical checks outlined above: the idea is to explore if
there is any systematic relation between the residuals and any one of the
explanatory variables. White's test and the Glejser test illustrated below
fall in this category. Let us have a look at each in turn.

Bartlett's test
Bartlett's test can be applied to check for the equality of the variances of
the dependent variable across groups defined by an explanatory variable.
Dealing with heteroscedasticity 257
The conditional variance of Y given X is the same as the conditional vari-
ance of the error term, u}. Indeed, using equation (7.1), we get:
V(Y) = V(l3 1 + 132 X+ Ex)

but, since X is given, Ex is uncorrelated with X, i.e. Cov( EX' X) = O, and


the variance of (13 1 + 132 X) is zero for any given X, we have (see Appendix
2.1 in Chapter 2):
V(Y) = V(l3 1 + 13 2 X) + V(Ex) = V(Ex) = u}
Hence, one way of checking for heteroscedasticity is to test for the
stability of the conditional variance of Y across the range of X in the sample
data. In practica! situations, we generally do not have multiple observations
of Y for a given X. The application of Bartlett's test, therefore, involves
that we first sort the data in ascending arder of the explanatory variable
which is suspected to be the cause of the heteroscedastic pattern of the error
term, and divide the sample in to several groups, say k groups (classes ),
based on this explanatory variable, after which we subsequently test the
hypothesis of homogeneous variances across the groups.
If u? is the variance of Y in the ith class, the null hypothesis we seek
to test can then be stated as

Now let Y;j = jth Y value in the ith class; n; = the number of observations
in ith class; and f; = (n; - 1); f = "'2.f;. The test is then performed as follows:

1 Compute the sample variance for each group i:


1 n -
s.2 = - - (Y. -
, n- _ 1 '('i
~ ,1
Y.)2
,
l 1=1

where
- 1 n
Y;=-~ Y;j
n; j=l

s? is the estimator of u¡2, i = 1, 2, 3, ... , k.

2 Compute the pooled sample variance of all the groups together:


k k
_L f;s/ _L f;s/
s2 = i=l _ i=l
k f
_Lf;
i=l

s2 is the estimator of u 2 under H 0 •


258 Econometrics for developing countries
3 Under the null hypothesis the ratio AIB has approximately a chi-square
distribution with (k - 1) degrees of freedom, where:
k
A= f· ln (s2 ) - LÍi' ln (s/)
i=l
(7.3)
B=l+ 1 [L G)- --flJ
3(k - 1)
k
i=l i

In our example, we first sort the sample of 261 workers in ascending


order of age, divide it into three groups of 87 workers each, and compute
the means and sample variance s;2 of weekly wage income for each group
(these calculations are summarised in Table 7.1). The computation of s2,
A and B are all calculated in accordance with the formulae stated above.
The larger the computed value of the statistic A/B, the more suspect the
null hypothesis of homogeneity of variances across the three classes will
be. Thus, the test consists in comparing the computed chi-square value
with the 5 per cent or 1 per cent critical value of the chi-square distrib-
ution for the corresponding degrees of freedom (2 = 3 - 1). The null
hypothesis is rejected if the ratio AIB exceeds the critical value. Most
statistical software provide the means to use Bartlett's test, or it can be
done on spreadsheet (see, for example, Table 7.1). The test statistic
computed for wage income was 47.8 which is significant at 5 per cent level.
Consequently, we reject the null hypothesis.
Note, however, that Bartlett's test is only valid under the assumption
that the dependent variable is distributed normally, which is not the case
with wage income. Not surprisingly, the distribution of weekly wage
income data is highly skewed. Bartlett's test, therefore, is not valid for
this reason. But its utility líes in exploring transformations which can

Table 7.1 Calculation of Bartlett's test for income data


Sample /neo me Logged income
Mean 1 88.0 4.2
2 179.6 4.9
3 232.1 5.2
s.2l 1 4734 0.49
2 16676 0.52
3 22096 0.61
ln(s¡2) 1 8.46 -0.72
2 9.72 -0.66
3 10.00 -0.50
sz 14502 0.54
ln(s2) 9.58203 -0.62223
A 48.05 1.15
B 1.01 1.01
Bartlett's 47.81 1.15
Dealing with heteroscedasticity 259
remove heteroscedasticity. Following the ladder of power transforma-
tions discussed in Chapter 3, we perform the logarithmic transformation
of the weekly wage income, and the transformed distribution turns out
to be approximately symmetrical according to the exploratory checks
illustrated in Chapter 3. The skewness-kurtosis test for normality of
the logarithms of wage income also turns out to be insignificant at the
5 per cent level. Hence, we can apply Bartlett's test to the transformed
data (log weekly wage income ), following the same steps explained above.
The only difference this time is that we compute the sample variances
of log income for the different groups instead of those of income
itself. The computed test statistic now works out to be 1.15, which is
insignificant at the 5 per cent level. The test is valid this time as the
hypothesis of normality of the distribution of log income could not be
rejected by the skewness-kurtosis test. We can, therefore, conclude that
log transformation successfully removes the heteroscedasticity observed
in the raw data.
How do we divide the observations across the different groupings? In
this example, arriving at the number 87 to divide the sample into three
groups was easy because 261 is divisible by 3. If the sample size is not
divisible by 3, take the next highest integer of the quotient, say n, and
divide the sample so that the bottom and the top groups have n units
each, and put the remainder of the data in the middle group. For example,
if the sample size were 250, then 250/3 = 83.33 is rounded to 84, and the
sample could be divided into 84 + 82 + 84. Here, we used three groups
as a matter of convenience. The point is to choose a suitable number of
groups (minimum of two) depending on the total sample size. Too many
groups with too few data points in each groups will not lead to reliable
conclusions. If heteroscedasticity is present and increasing (or decreasing)
with one of the explanatory variables, this will show up if you use a few
(preferably three or more) sizeable groupings. Two groupings is not so
advisable because they border each other and, hence, they will have similar
variation where they meet. Only use two groups when your sample size
is rather small.

Goldfeld-Quandt test
Bartlett's test was used to check for homogeneity in the conditional vari-
ances of Y, the dependent variable. The Goldfeld-Quandt test checks for
homogeneity in the conditional error variances. Hence, with Bartlett's test,
you group Y with reference to the ascending order of one of the X vari-
ables, but you do not run groupwise regressions. The test is performed using
the computed conditional variances of Y. As we shall see, the Goldfeld-
Quandt test also implies that you group the data with reference to the order
of one of the X variables, but in this case you run groupwise regressions
to obtain sets of within-group residuals. The test is commonly used
260 Econometrics far developing countries
when the heteroscedastic variance a 2x is suspected to vary monotonically
(i.e. consistently increasing or decreasing) with one of the explanatory
variables in the regression model. The procedure is based on dividing
the sample into three groups in ascending arder of one of the explanatory
variables, and testing for the difference in the error variance between the
bottom and the top groups. Hence, the middle group is not considered in
the test. Its only function is to prevent the extreme groups bordering on
each other.
The test involves the following steps:
1 Arrange the data in ascending arder of the explanatory variable
suspected to be related to the error variance.
2 Drop a number of the middle observations, say e, so that (n - e) is
divisible by 2, hence n' = (n - c)/2 is the subsample size. A rule of
thumb is to drop about 1/4 of the total observations from the middle.
3 Estimate two separate regressions for the bottom and the top group
of observations, and compute the corresponding residual sums of
squares - respectively, RSS1 and RSS2.
4 Compute the ratio of the higher to the lower residual sums of squares.
This ratio has an F-distribution with [d, d] degrees of freedom where
n-c
d = - - - k = n' - k
2
k = No. of estimated coefficients

under the hypothesis that the error distribution within each group
is normal, and that the error variances are the same. The higher the
computed ratio, the less likely it is for the hypothesis to be true.
5 Compare the computed ratio with the critical value of the relevant
F-distribution. If the computed exceeds the critical value, then the
hypothesis of homoscedasticity is rejected.
Let us use our sample data of 261 workers again to illustrate this test step
by step:
1 Since there is only one explanatory variable, age, we arrange the data
in ascending arder of age.
2 The total number of observations is 261. One-fourth of 261 is 65.25.
Now (261 - 65) is an even number and, hence, we drop the middle
65 observations. This leaves us with 98 (i.e. (261 - 65)/2) observa-
tions each in the bottom and top groups. The bottom group corre-
sponds to lower values of the age variable and the top group to the
higher values.
3 Bottom group (observations 1 to 98): RSS1 = 382,302
Top group (observations 164 to 261): RSS2 = 2,207,120
4 Fcalculated = RSS2IRSS1 = 5.63.
Dealing with heteroscedasticity 261
5 The computed value of 5.63 is higher than the critica! value in the
F-distribution with [96, 96] degrees of freedom at 5 per cent level
of significance. Hence, we reject the null hypothesis of homosced-
asticity.
Again, this test is not strictly valid because of the non-normality of the
distribution of wage income. If we repeat this test with log income, this
time the test is valid, and the corresponding calculated value much lower
(1.96), though not insignificant.
It is not necessarily the case that the different tests for heteroscedas-
ticity will lead to the same conclusion. These tests depend on the way in
which we divide the data into groups, hence, different groupings may yield
different results. We shall come back to this point later.

White's test
The basis for this test is to check whether there is any systematic relation
between the squared residuals and the explanatory variables. This is
achieved by regressing the squared residuals e;2 on all the explanatory
variables and on their squares and cross products. Thus, if X 1 and X 2 are
the explanatory variables, then White's test involves regressing e2 on X 1,
X 2 , X 12 , Xz2 and X 1 .X2 , and using the overall F-test to check if the regres-
sion is significant or not. This test (and others like it) is in fact a general
regression specification error test (RESET) and not solely a test for
heteroscedastic errors (see Box 7.1).
In our example of the regression of wage income on the age of the
worker, the only explanatory variable, we regress the squared residuals
e2 from the fitted regression in (7.2) on AGE and AGE2, which yields an
F-statistic of 6.58. This calculated value should be compared with the crit-
ica! value F(z,zss)· The calculated F-statistic is significant at the 1 per cent
level and, hence, we reject the hypothesis that there is no relationship
between squared residuals and age and age squared. By implication, we
reject the hypothesis of homoscedasticity.

Glejser's test
Like White's test, Glejser's test also checks whether a systematic relation
exists between the residuals and the explanatory variables. However,
Glejser approaches the problem in a different way. The test involves
regressing absolute residuals separately on X, x-1 and X1 12 , and uses
t-tests for the slope coefficients to be zero. If there is more than one
explanatory variable then this exercise is to be repeated for each of the
explanatory variables. The hypothesis of homoscedasticity is rejected
if any of the slope coefficients turns out to be significantly different
from zero. The difference with White's test, therefore, is that Glejser's
262 Econometrics far developing countries

Box 7.1 White's test as RESET (regression specification error test)


The presence of heteroscedastic residuals can result from hetero-
scedastic error terms in a model which is otherwise correctly spec-
ified. However, heteroscedastic residuals can also occur because of
misspecification of the regression model itself: in particular, because
one or more relevant variables have been omitted from the regres-
sion or because of non-linearity of the regression curve.

Heteroscedastic residuals due to omitted variables


If a variable has been incorrectly omitted from a regression model,
the residuals of the estimated model will incorporate the effect of
this omitted variable. In so far as the omitted variable varies (in
whichever way) with one or more of the regressors included in the
model, heteroscedastic residuals may result. For example, if an
omitted variable Z varies strongly with regressor X, the variation in
the observed residuals is likely to reflect this covariation (whatever
its form: linear, quadratic, etc.) between Z and X. White's test, which
involves regressing squared residuals on the regressors, their squares
and their cross-products, is likely to pick up this effect of omitting
a relevant variable.

Heteroscedasticity and non-linearity


Similarly, misspecification due to non-linearity can lead to hetero-
scedastic errors. For example, in a simple regression of Y on X,
fitting a straight line to a scatter which reflects a typical quadratic
shape will produce U-shaped residuals, the variation in which is
determined by X.
White's test, therefore, is rather indiscriminate in its scope. It tells
you when things are wrong but <loes not necessarily point the direc-
tion to look for the cause. In this sense, it is a general misspecifica-
tion test which prompts you to see what the problem could be. It <loes
not allow you, however, to jump toan immediate conclusion.

test selectively investigates which explanatory variable accounts for the


heteroscedasticity as well as the form it takes. White's test, in contrast, is
akin to a broad-spectrum diagnostic device aimed at checking whether
something is wrong. Let us try out Glejser's tests on the income-age data.
The results are as follows (standard errors in parentheses):
/\
le 1= 4.52 + 2.55 AGE
(0.39)
Table 7.2 Summary of tests for heteroscedasticity with values for various regressions
Statistic Definition Test /neo me Log income Log income
statistic on age on age on age, education
and sex

k
L ft' In
f · In (s2) -

1+ . 1 [kG)
i=l

2.: - --~lJ
Bartlett's 3(k 1) f í=l i x2 47.81 1.15 Age: 1.15
Edu: 1.77
RSS1
Goldfeld-Quandt RSS2 F 5.63 1.96 1.42

White's F-test on null hypothesis that R2 == O in


regression of squared residuals on regressors,
their squares and cross-products F 6.58 1.93 1.47ª

Glejserb Test of significance of slope coefficient from Age: 6.5 Age: 1.9 Age: 1.8
separate regressions of absolute residual on 1/Age: 6.4 l!Age: 1.4 1/Age: 0.8
regressor, its inverse and its square root .,/Age: 4.5 JAge: 1.9 ,/Age: 1.6
Notes:
ª Sex squared omitted from \Vhite's test for multiple regressíon owing to multicollinearity in the test equation.
b t-statistics for Glejser test are absolute values.
264 Econometrics for developing countries
A 1
le 1=163.97 - 2159.0 AGE
(335.5)

I~ 1= -81.74 + 30.18 VAGE


(6.69)

The two-sided t-tests for the slope coefficients equal to zero are signifi-
cant at the 1 per cent level in all three regressions above (see Table 7.2).
Thus, Glejser's test also rejects the hypothesis of homoscedasticity.
lt is quite possible that not all the four tests - Bartlett's, Goldfeld-
Quandt, White's and Glejser's - produce the same results in terms of
either rejecting or accepting the hypothesis of homoscedasticity. If all the
tests fail to reject homoscedasticity then we are on firm ground to proceed
with the initial model. If at least one test rejects homoscedasticity then
we should examine carefully the nature of heteroscedasticity by means of
the graphs discussed earlier and proceed according to the principles which
we shall discuss in the next section.

Exercise 7.3
Using the cases listed in exercise 7.1, try out all four tests discussed in
this section to test for heteroscedasticity.

Exercise 7.4
Continuing with exercise 7.2, do the tests for heteroscedasticity with
the model featuring the logarithms of income versus the age of Indian
workers.

7.4 TRANSFORMATIONS TOWARDS HOMOSCEDASTICITY


Heteroscedasticity of the error term implies that we cannot use the esti-
mates of the standard errors of the regression coefficients computed on
the basis of the standard formulae derived from the classical regression
model. Furthermore, the least squares estimators are inefficient, but un-
biased. Consider, for example, our income-age relationship: we can use
the estimated regression (7.2) for predicting the average weekly wage
income of the workers of a given age without falling prey to a statistical
bias. But this prediction is not the most reliable (since it is inefficient)
among all linear predictions and we cannot make any statement about
the uncertainty (confidence interval, hypothesis tests) of the predictions
based on the standard errors computed according to the formulae under
the assumption of homoscedasticity.
Why not modify the formulae accordingly to allow for heteroscedastic
errors? In principle, this is possible. But to do so we need to be able to
Dealing with heteroscedasticity 265
state the nature of the heteroscedasticity of the error term. This is exactly
what we try to do, but not necessarily by forsaking the advantages of the
normal linear regression model. As we shall now see, transformations
often do the trick by eliminating heteroscedasticity in the residuals, indi-
cating that our model may well have had the wrong functional form.
Indeed, we find that in very many situations in practice transformations
help us to get rid of heteroscedasticity.

Finding an appropriate transformation


One of the most common reasons for heteroscedasticity is the skewness
of the distribution of one or more variables involved in a regression model
with socioeconomic data. In such cases, a power transformation - and, in
particular, the logarithmic transformation - often eliminates the skewness
in the data. How do we find the appropriate transformation to eliminate
heteroscedasticity? As we have seen, heteroscedasticity implies that the
conditional variance of the dependent variable Y or the error term E varies
with the conditional mean of Y. Now, 'if the functional relationship
between the variance and the mean is known, a transformation exists
which will make the variance (approximately) constant' (Rawlings, 1988:
309); and one such general functional relationship between the conditional
variance of the dependent variable and its conditional mean is commonly
applied in practice which allows us to use power transformations to elim-
inate heteroscedasticity. That is, if:
(7.4)
or, alternatively:
ay= Aµ} (7.5)
which we can conveniently re-express as a double-log equation as follows:
ln(cry) =A+ kln(µy) (7.6)
where, cry2 = the conditional variance of Y; µy= the conditional mean of
Y; and, A and k are constants. That is, if the standard deviation of Y is
proportional to sorne power k of its conditional mean, then the power
transformation Y1-k when k is not unity and the log transformation when
k = 1, will (approximately) eliminate heteroscedasticity (ibid.). Thus,
if k = 1 and, hence, the standard deviation is proportional to the mean,
the logarithmic transformation will do. If, however, the variance (and not
the standard deviation) is proportional to the mean and, hence, 2k = 1,
the square root transformation will do the trick.
But, typically, cry and µy are unknown. So how do we find the appro-
priate value for k? What we can do is to substitute the absolute residuals
1 e for ay and the predicted YP for µy in equation (7.6) and use the data
1

to estimate its slope coefficient, k, with least squares. In our example of


266 Econometrics far developing countries
6.56785 o

o o 8
o
§ º§ oºo o o
Q) ºº0 o s8o o o
Ol o 0
8ofil~o @º§@o@
Cll
e
o o8ºog gº 8 o8
o o §o ºª8º0 ºº g
(]) o o o g§ ~ 0
o @º oo 0
° 0
o o
o o
E g o o o g o o ºº o o o
8 º º 8 º º eºe º º º o
.!: o 8 oºº o ºº o
¡,; ooº ºooºso o o
Cii o o o o0 o o
::i
o o
"O o o
·¡¡; @~§ o
o
~ o
o
o
Ol o o o
.3 o o
o

o
o o
-0.158628
4.13298 5.92665
Lag (predicted wage income)
srara™

Figure 7.4 Log j e versus predicted Y


1

the income-age data the corresponding scatter plot with regression line
is depicted in Figure 7.4.
The plot indicates a linear relationship between ln(jej) and the ln(Yp),
where YP is the predicted Y. The corresponding fitted regression is given
as follows:
ln(j e j) = -2.0930 + 1.2108 ln(Yp)
(0.1691)
which reveals that the slope coefficient is reasonably close to unity.

Exercise 7.5
Test formally whether the slope coefficient in equation (7.7) is significantly
different from one, using the 5 per cent significance level. What do you
conclude from this test?

You will have found from the two-sided t-test that the slope coefficient is
insignificantly different from one at the 5 per cent significance level.
Hence, it seems a good idea to try out the logarithmic transformation to
see whether it helps us to eliminate heteroscedasticity. The regression of
ln(Y) on AGE yields the following results (t-statistics in brackets):
Dealing with heteroscedasticity 267
2.61322 o

o
o o o
o
o o o
o 8
o o
o o o o
o o o o o
o o o o o
ºo o g
o o o 8 o
o ºg oº o o o ºº o og o
o ºoºº o o 0
00°~ o o o o
o o o o
o o 8ºº o
0
§ o !!¡¡ 0 o
0 0°08 oººB 8 8 o o !! o
o o ºog o o o o o o
o o
OOOO!! § § 8 O!! o o
o0 o o o 0° o o
o oo ~ oº o o ~o 0 o o 800°
o o oº ~8 ºosº o o
ºo o o§gº o 90 O ¡¡oºº oºo ºº o o
0.003345 o ºoo!! o o o 0 o O

4.08787 6.21782
Predicted log wage income
STaTa""

Figure 7.5 Scatter plot of absolute residuals versus predicted

ln(Y) = 3.72 + 0.03AGE R 2 = 0.23


(28.9) (8.89)
Figure 7.5 plots the absolute residuals of this regression against its
predicted log income. As you can see, this scatter plot no longer shows
a heteroscedastic pattem and, hence, our transformation proved to be
quite successful.
The plots (not shown here) of the residuals and the squared residuals
against the predicted values also confirm that there is no longer any
systematic pattem in the scatter. This should not surprise us: our earlier
application of Bartlett's test to the log transformed income data led us to
conclude that the hypothesis of homogeneous variances of the log income
data is acceptable. All other tests (Goldfeld-Quandt test, White's test and
Glejser's test) come to a more or less similar conclusion. We can now be
reasonably confident, therefore, that our transformation resolved the
problem of heteroscedasticity.

Exercise 7.6
Using the data in INDIA, perform Bartlett's, Goldfeld-Quandt, Glejser's
and White's test for the regression of log income on age. Comment on
your results.
268 Econometrics far developing countries
1800 o
o

o
o o
o o
o
o
~
.a o
'6
e o
Q)
a. o
><
Q)
""O
o
o
u.

83 o

382 3832
Total expenditure

Figure 7.6 Food and total expenditure

7.49554 o o
o
o o o
o o

o
o
o

~
.a
'6
e
Q)
a.
><
Q)
""O
o o
.E o
Ol
o o o
...J
o

4.41884 o

5.94542 8.25114
Log total expenditure

Figure 7. 7 Effect of log transformation


Dealing with heteroscedasticity 269
The results you should achieve are summarised in Table 7.2 Compared
with the regression of untransformed income on age, a dramatic reduc-
tion in each of the calculated test statistics can be observed. It therefore
does seem, as observed in the previous paragraph, that the log transforma-
tion of income has removed the problem of heteroscedasticity (though we
return to this point below).

The versatility of the logarithmic transformation


Once more, the logarithmic transformation has shown its versatility to
rectify problems to do with skewed distributions. Let us conclude this
section with a final graphical example of the use of logarithms.
Figure 7.6 presents the scatter plot of household food expenditure
against total expenditure. The data (in the file INDFOOD) were obtained
from a household sample survey of a selection of villages in India. The
sample size is 217 households. The skewness in both the distributions is
clearly evident from the scatter plot. So is the inherent pattern of
heteroscedasticity, although a straight line seems to be an appropriate
description of the average relationship. The effect of log transformation
of both the variables is shown in Figure 7.7. As you can see, skewness
has been removed from both these distributions and the scatter plot no
longer indicates any sign of heteroscedasticity.

Heteroscedasticity as a symptom of omitted variable bias


In the example so far we have examined the impact of age on earnings.
But after our analysis in Chapter 6, we may well expect that this equa-
tion suffers from misspecification due to omitted variables. Surely other
factors, such as education, affect earnings. The data file INDIA also
includes data on the age and sex (a dummy variable) of the workers.

Exercise 7.7
Using the data in INDIA, regress log income on age, education and sex.
Test the hypothesis that education and sex may be dropped from the equa-
tion. Perform Bartlett's, Goldfeld-Quandt, Glejser and White's tests.
Comment on your results.

The results you should achieve for exercise 7.7 are shown in Table 7.2.
All the calculated test statistics, except Bartlett's, have fallen still further
and the null hypothesis may now be accepted in the case of the Goldfeld-
Quandt test. This example shows that residual heteroscedasticity may well
result from omitted variable bias, and the inclusion of incorrectly omitted
variables will reduce evidence of heteroscedasticity.
270 Econometrics far developing countries
The statistic for Bartlett's test did not change since it is calculated with
reference to the dependent variable. It is therefore unchanged by changes
in the functional form of the regressor rather than regressand. Except
White's test, all the tests are calculated with respect to a single regressor
(age in the example here). If there are more regressors, then the data
should be sorted by each regressor in turn and the test recalculated for
Bartlett's (as shown in Table 7.2) and the Goldfeld-Quandt tests, and the
absolute residual regressed on each regressor (and transformations
thereof) in turn for the Glejser test. These steps are not necessary if there
is sorne good reason to believe that the heteroscedasticity is related to
one specific regressor.

7.5 DEALING WITH GENUINE HETEROSCEDASTICITY:


WEIGHTED LEAST SQUARES AND HETEROSCEDASTIC
STANDARD ERRORS
The previous section showed that the presence of heteroscedasticity in
the residuals is probably a result of model misspecification. The example
we used of Indian workers showed how changing functional form and
adding incorrectly omitted variables reduced the evidence of hetero-
scedasticity. However, it can also happen that heteroscedasticity remains
a problem even when we are sure that the model is correctly specified,
which could be the case for the consumption function in which richer
people enjoy more variation in their consumption than those who are
less well off. In such cases we must try to cure the problem. The tradi-
tional solution is weighted least squares, which permits BLUE estimates
to be obtained from ordinary least squares. However, weighted least
squares is not always appropriate, in which case we may use
heteroscedastic consistent standard errors. We consider each of these
possibilities in turn.
There are occasions when we can model the heteroscedastic error vari-
ance of the population regression (7.1) in the following form:
ay= w? a2 (7.9)

where i = 1, ... , n, the sample observations; a¡ = the error variance for


observation i; W¡ = the weight for observation i; a = a constant variance.
Hence, the specific error variance for observation i can be obtained by
multiplying a constant variance, a, by the corresponding weight of obser-
vation i. As we shall see, if the set of weights are known or if their values
can be easily hypothesised, it is possible to obtain efficient estimators of
the coefficients of the regression model and make valid statistical infer-
ences. A typical case where we can do so is when we have grouped data.
So let us first consider this case.
Dealing with heteroscedasticity 271
Table 7.3 Grouped household expenditure
Per capita average Per capita average Per capita average
expenditure class total expenditure food expenditures
Below 175 145.67 83.39
175-225 199.42 115.96
225-275 247.36 143.95
275-350 301.92 165.86
Above 350 437.93 206.70

Working with grouped data


In development research we often deal with published data which are
typically presented in the tabular form of group means. Table 7.2 is an
example of such data. The table lists per capita total and food expendi-
tures by per capita expenditure groups for the data depicted in Figure 7.6.
Suppose the correct model is given as follows:
(7.10)
where E is homoscedastic with variance a 2 • This model applies to the data
on individual households. However, the data as listed in Table 7.3 are
grouped and, hence, model specification (7.10) will not be appropriate. In
general, if there are k grouped means, then the appropriate model for the
data can be written as follows:
- -
Y; = 131 + l32X; + E; i = 1,2,3, ... ,k (7.11)

where the bars on the variables Y, X and E denote the corresponding


group means.
But, in this case, due to the grouping of the data, we get:

V(E;) = a2 i = 1,2,3, ... , k (7.12)


n;

and, hence, the model for grouped data has heteroscedastic error terms.
But note that the error variance is of the form as specified in (7.9) with
w? = (l/n;).

Weighted least squares


When the weights are known, as in the case of grouped data, a simple
transformation of the variables in the model provides us with a way to
apply the classical model. In general, if the heteroscedastic model is given
as follows:
(7.13)
272 Econometrics far developing countries
where V(E;) =u?= w? u2, we can divide equation 7.13 by W; so as to get
the fallowing model:
y 1 X E
___!_ = f31 - + f32 _l + _l_ (7.14)
W¡ W¡ W¡ W¡

This transfarmed regression has no constant term. The variance of the


error term in the new specification is homoscedastic because

V (5_) = J_
W;
V( E.) = J_ w. u2 = u2
w/ ' w;2 '
2
(7.15)

Therefare, if we regress (Y/w;) on (llw;) and (X/w;) without a constant


term, the least squares estimators of 13 1 and 13 2 will have the BLUE prop-
erties. This procedure of estimation of the regression coefficients subject
to the additional assumption (7.9) is called the weighted least squares
method of estimation.
So much far the method in theoretical terms. The crucial issue in prac-
tice is to check whether the specification (7.9) of the error variance is
correct and to find the appropriate weights, w;. If we are dealing with
grouped data such as shown in Table 7.3, then W; = l!n¡. However, we
shall have to be reasonably sure that the regression model (7.10) is valid
in arder to use model (7.11). This is important. lf, far example, the regres-
sion of per capita faod expenditure on total expenditure has a double-log
farm, then the model (7.11) can no longer be validly applied to the
grouped data since the group averages are calculated from the original
values and not the log-transfarmed values.

Exercise 7.8
Using the data in the file TPEASANT (farm size and household size in
Tanzania), estímate the regression of landholding size on household size
with weighted least squares. Do you think that the resulting regression
satisfies the assumptions of classical linear regression?

When X¡ is an appropriate weight


There are other situations where assumption (7.9) seems to be appropriate.
One special case is when the X variable can serve as the relevant weight:
(7.16)
and, hence:
W¡ = X; (7.17)
which states that the error variance is proportional to the square of (one
of) the explanatory variable(s). This case is commonly applicable. With
economic data, far example, the variance of Y often tends to increase
Dealing with heteroscedasticity 273
with X. If the relationship of the variance with X can be adequately
described by equation (7.16), then we can use a transformed regression
equation specified as follows:
Y;_ 1 1
- - f31 - + f32 + E¡ (7.18)
X; X;
The interesting aspect of this specification is that the intercept and slope
coefficients in the model (7.13) have interchanged their places in the new
specification (7.18): ¡3 1 is now the slope coefficient in the transformed
equation and ¡3 2 is its intercept. Let us try this special case with the exam-
ple of the data on household total expenditure and food expenditure used
in Figure 7.6. We showed in Figure 7.7 that the logarithmic transformation
of both food and total expenditure removes the heteroscedasticity.
Therefore, the double log transformation gives us one way to model the
data. But let us try a different avenue of modelling the same data.
To do this, we examine the residuals from the fitted regression of food
(F) on total expenditures (1). The estimated regression is given below
(standard errors in brackets) as:
F = 88.51 + 0.47T (7.19)
(21.61) (0.02)
The standard errors are not usable due to heteroscedasticity, as is
evident in Figure 7.8 which plots the residuals against the predicted values.
Figure 7.9 present the plots of the absolute residuals of regression (7.19)
against total household expenditures. A linear relationship between e 1 1

and T is discernible in Figure 7.9, which implies that a quadratic rela-


tionship between e2 and T2 seems plausible. In other words, it seems that
assumption (7.9) is worth a try in this case.
The estimated transformed regression is:
~
(F!T) = 0.50 + 56.49 (111)
(7.20)
(0.02) (17.80)
Figure 7.10 presents the scatter plot of the absolute residuals of this
regression against (1/1). The pattern that we observed in Figure 7.9 has
now disappeared in Figure 7.10, indicating that the error term of the trans-
formed model is homoscedastic. The other two plots of the raw residuals
and the squared residuals (not shown here) also show the same absence
of any heteroscedastic pattern. The tests for heteroscedasticity all yield
insignificant results. Exploratory checks as well as the skewness- kurtosis
test for normality indicate that there is no evidence of deviation from the
normal distribution of the error term of the transformed model. We can,
therefore, accept the transformed model as adequate in the sense that the
standard assumptions of the classical normal linear regression model hold
in the case of the transformed model.
274 Econometrics for developing countries

o o
o
~
.a o
o
o
'O o
e:
Q)
a.
X
Q) o
"O o
o
E o
ro::;¡
"O
·¡¡; o
Q) o
a: o

o o o

269.685 1905.96
Predicted food expenditure
s ta 1 a•M

Figure 7.8 Residual versus predicted plot

582.7 o

o o

ro::;¡ o
"O
o
·¡¡; o
o
~ o
Q) o
"5 o o
o(/) o o
.o oºoo oºº o
<! ~o o '2P o
0
O (jo CD & O ~O o
~ ºº o o o o
0 'O~oO~ ;50 00 o o
o o o cP
o ~o 8 'b '(, c8 o
8=>
000 o
º~§:¡
o
.Glm.~
8 ºóJV~ o oo o o o
g' g:¡scg 'ti> o o o
0.608848
382 3832
Total expenditure
STaTa"'

Figure 7.9 Absolute residuals: regression of Fon T


Dealing with heteroscedasticity 275

0.475001 o

o o
o
o o o
o o o
o

o
o
o o

o o
o o o o o

o o o
0.000069
0.000261 0.002618
Inversa total expenditure

Figure 7.10 Absolute residuals: regression of FIT on l/T

1989.55

Q)

~
ue:
g¡_
X
Q)
"O
o
o
LL

83
382 3832
Total expenditure
STaTa"'

Figure 7.11 Comparing double-log and linear fitted regressions


276 Econometrics far developing countries
Let us now compare the estimates we obtained from regressions (7.19)
and (7.20), befare and after transformation. The estimates of 13 2 are fairly
clase: 0.47 and 0.50. But, the estimates of 131 differ substantially: 88.51 and
56.49. The estimates of the standard error of the coefficients are not very
different but higher in the transformed model. Given that the transformed
model conforms to the assumptions of the classical normal linear regres-
sion model, the latter estimates can be used to make inferences from the
sample data. The R 2 values are not comparable between (7.19) and (7.20)
which explains why we did not bother to report them.
This leaves us with an interesting question. We have seen that the
double-log model allows us to model these data in an acceptable way.
Similarly, the linear model, suitably adapted to deal with heteroscedastic
errors, also gives us a viable way to model the same data. So which model
should we use? This is impossible to sort out, at least in the present state
of statistical theory. The point is that we are resorting to ad hoc trans-
formations to find an adequate description of the data. The problem arises
because two different models may be equally data admissible. You may
find this surprising, but the reason is easy to see if, as shown in Figure
7.11, you compare the fitted regression line and the curve produced by
both models. On statistical grounds, therefore, both models perform
equally satisfactorily and, hence, there is no reason to prefer one to
the other. But we might decide on substantive grounds. The double-log
specification gives us a constant elasticity of food expenditure with respect
to the total expenditure whereas the slope coefficient in the linear
model (which is the intercept in regression (7.20)) is a marginal propen-
sity. Hence, much depends on which measure you find more appropriate
to use.

White's heteroscedastic consistent standard errors (HCSEs)


Whilst weighted least squares allows us to obtain BLUE estimates from
ordinary least squares we have seen that it is not without its problems.
Specifically, we need a model of the heteroscedasticity, and even then the
procedure may not be appropriate if the correct functional form of the
regression differs from the functional form of the relationship between
regressor and error term. For these reasons much current practice favours
the use of heteroscedastic consistent standard errors.
In the two-variable case we saw that:

Var (b2) = L (X¡ - X)2 (7.21)

where the formula is simplified using the assumption that E( E¡2) = u 2 for
all i. Where this assumption is not valid (i.e. the errors are heteroscedastic)
then:
Dealing with heteroscedasticity 277
~ (X. - X)2 (T2
Var (b 2) = ~(i; _X) 2 ' (7.22)

White (1980) showed that substituting the squared residuals (e¡2) into
equation (7.22) yields a consistent estimate of the standard errors (this
result generalises to the k variable case). However, unlike with weighted
least squares, these are not the minimum variances.
Inspection of equation (7.22) shows that if the errors are homoscedastic
then the expression simplifies to that in equation (7.21). That is, the het-
eroscedastic consistent standard errors and those usually reported will be
the same if there is no heteroscedasticity. A divergence between these
two sets of standard errors is thus a rough test for the presence of
heteroscedasticity.

7.6 SUMMARY OF MAIN POINTS


1 Heteroscedasticity is a non-constant error variance across the sample.
The presence of heteroscedasticity renders least squares estimators
inefficient, but they remain unbiased. In other words, they are linear
unbiased estimators, not best linear unbiased estimators. Moreover,
the standard formulae for the standard errors of the coefficients no
longer apply and, hence, statistical inferences based on the t-test or F-
test are not valid.
2 To detect heteroscedasticity we can use visual diagnostic tools: more
specifically, we use the plots of raw, absolute and squared residuals
against the predicted values of the dependent variable or against each
regressor in turn.
3 Tests for heteroscedasticity fall into two categories. The first involves
grouping the data with respect to different ranges of one of the
explanatory variables and testing whether the conditional variances of
the dependent variable or of the error term are the same. The second
approach tests whether any systematic relation exists between the error
term and any of the explanatory variables. Bartlett's and Goldfeld-
Quandt's tests are examples of the former, while White's and Glejser's
tests apply the latter principle.
4 White's test also serves as a general regression specification error test
(RESET). As such, it is a broad-spectrum diagnostic test, which tells
you when something is wrong but not necessarily what is wrong. The
problem can be heteroscedastic errors, omitted variables or non-
linearity of the regression curve.
5 Heteroscedasticity may be detected as a symptom of model misspeci-
fication (either incorrect functional form or omitted variables) rather
than a genuine problem in the residuals of the true model. Hence
model respecification is the first course of action to take in the pres-
ence of heteroscedastic residuals.
278 Econometrics far developing countries
6 Power transfarmations often remove the problem of heteroscedasticity.
A plot of the logarithms of the absolute residuals against the loga-
rithms of the predicted values of the dependent variable allows us to
judge whether a transfarmation is likely to salve the problem of
heteroscedasticity. The slope coefficient of the corresponding regres-
sion tells us which transfarmation may be most appropriate.
7 If we believe we are dealing with a case of genuine heteroscedasticity
then, in sorne cases, the method of weighted least squares allows us
to derive efficient estimators of a regression model with heteroscedastic
errors and to make valid inferences. Two cases are common in this
respect: regressions with grouped data and situations where the error
variance varíes proportionally with one of the regressors.
8 If weighted least squares is not possible, then unbiased, though not
the most efficient, standard errors may be calculated using White's
heteroscedastic consistent standard errors (HCSEs).

ADDITIONAL EXERCISES

Exercise 7.9
Use the data in data file INDFOOD to test far heteroscedasticity in the
regression of household faod expenditure on total expenditure. Repeat
the tests using the lag of both variables. Comment on your findings.

Exercise 7.10
Using the data in data file LEACCESS, regress life expectancy on (a)
income per capita; (b) logged income per capita; and (c) logged income
per capita and access to health. Test far heteroscedasticity in each regres-
sion equation. Comment on your results.
8 Categories, counts and
measurements

8.1 INTRODUCTION
Categories matter a great deal in empirical analysis. The reason is that
the average level of a numerical variable or relations between variables
may differ quite markedly across different categories. For example, rural
or urban location affects consumption and production patterns of house-
holds. Similarly, wage and salary earnings may differ between men and
women, even for the same level of education and years of experience.
Occupational status affects both the health (for example, mortality rates)
and the wealth of people. Categorical variables allow us to classify our
data into a set of mutually exclusive categories with respect to sorne qual-
itative criterion: for example, men/women; rural/urban; occupation; region,
countries or continents; policy regimes. In practice, this type of variable
is inevitably discrete in nature inasmuch as we only consider a definite
(usually limited) number of categories. For example, the gender variable
has only two categories (male/female), while a variable on occupational
status usually distinguishes among eight to ten categories. The distinctive
nature of these variables, therefore, is that they do not measure anything,
but assign a quality to our data (i.e. they are qualitative not quantitative ).
Hence, we cannot compute an average for this type of variable, but we
can count (frequencies) how many observations in our data set fall in a
group defined by a qualitative categorical variable. This chapter deals with
ways in which these variables can be employed in empirical analysis to
look deeper into the structure of our data. In particular, this chapter shows
that the use of categorical variables helps us to guard against making
unwarranted generalisations based on the assumption that homogeneity
prevails when, in fact, we are lumping things together which should be
kept separate. We have already come across categorical variables in
Chapter 6, when we introduced dummy variables. In this chapter, we take
this analysis further and look at the categories behind the dummies.
Throughout this chapter we shall use one extended example concerning
the effects of education and gender on weekly wage earnings for the sam-
ple of 261 workers in an industrial town of southern India (data set INDIA)
280 Econometrics for developing countries
to illustrate our argument on categorical variables. Section 8.2 deals with
the analysis of the relation between a numerical dependent variable and a
categorical explanatory variable, which involves comparing averages
between categories and, hence, is a natural extension of the principle of
regression. We show that dummy variables can be used to depict categories
of a categorical variable, a technique which allows us to extend the reach
of regression analysis to <leal with qualitative explanatory variables. Section
8.3 shows how to analyse the association between two (or more) categori-
cal variables in the context of a contingency table. This technique allows us
to test whether two (or more) categorical variables are statistically inde-
pendent or not, using a chi-square statistic, thus laying the groundwork for
discussing the regression between a quantitative dependent variable and
two (or more) categorical variables in section 8.4. Here we meet again the
by now familiar concept of partial association which we carne across in
Chapter 5. But we also introduce you to interaction effects which result
when two or more categorical variables interact in unison to affect the out-
come of the dependent variable. As usual, the last section 8.5 summarises
the main points of this chapter. Perhaps you are wondering whether cer-
tain problems involve the use of a dependent qualitative (categorical) vari-
able. They do, but they will be dealt with in the next chapter.

8.2 REGRESSION ON A CATEGORICAL VARIABLE:


USING DUMMY VARIABLES
Take a look at Table 8.1. On the left, workers are classified by gender;
on the right, weekly eamings are tabulated by weekly wage income group-
ings. Gender is a typical nominal categorical variable inasmuch as it
defines two categories (and, hence, it is dichotomous) which are named
but imply no ordering. While the frequency distribution of weekly income
(in Indian rupees) is obtained by first defining income groups and then
counting the number of workers in the sample that fall into each group,
the frequency distribution of gender is obtained simply by counting the
number of male and female workers. Obviously, we cannot talk about the
mean or median of a categorical variable as we do about income. But, as
we shall see in this and the next chapter, we can still analyse the propor-
tions based on counts in various ways.
Categorical variables can also be based on a measurement variable. For
example, it is generally quite difficult, if not impossible, to get the exact
figure for the monthly or annual income of households in a survey. This
is particularly true of developing countries because of the informal nature
of various sources of income which are hardly recorded, officially or other-
wise. One can, however, manage to get a rough figure of the income of
a household by using a combination of direct and indirect means.
Naturally, it will not be always very meaningful to use such shaky income
figures in a regression and, hence, it may be more useful to categorise
Categories, counts and measurements 281
Table 8.1 Frequency distributions of a categorical variable and a measurement
variable
Distribution of workers by gender Distribution of workers by weekly wage
Gender Frequency Per cent lncome Frequency Per cent
Mal e 206 78.9 Up to 70 64 24.5
Female 55 21.1 70-150 94 36.0
151-300 54 20.7
Over 300 49 18.8
Total 261 100.0 Total 261 100.0

households into a few income categories such as low, middle and high
income earners. This gives rise to income as an ordinal categorical vari-
able. We use the term 'ordinal' because the three categories involve an
ordering. In a similar way, a rainfall variable classified by low, medium
and high rainfall is also an ordinal variable. Another such variable, and
one we shall use in this example, is educational achievement: below
primary, primary, secondary and higher education. The latter example
shows that an ordinal variable is not necessarily derived from a measure-
ment variable. Years of schooling is a quantitative variable but educational
achievement is not exactly the same: it indicates that certain standards
have been reached and successfully accomplished.

Exercise 8.1
Using the data set SOCECON, construct an ordinal income variable by
grouping developing countries into low, lower-middle and upper-middle
income countries as measured by GNP per capita. To do so, use the
following cut-off points: low income countries, $600 or below; lower-
middle income countries: above $600 and up to $2,500; upper-middle
income countries: above $2,500 and up to $9,000. (Exercise 8.4 will require
you to analyse life expectancy by these categories, so be sure to sort these
data by the income categories when doing this exercise.)
Do men earn more than women? To answer this question with our data
we need to analyse the association between weekly earnings, a quantita-
tive variable, and a dichotomous qualitative variable. To do this, we need
to tabulate earnings by gender. This is done in Table 8.2. The income
differences between male and female workers come out quite sharply from
these two columns: almost 78 per cent of the female workers earned less
than Rs 150 while the corresponding proportion among their male coun-
terparts was about 56 per cent. These differences are reftected in the
average earnings of male and female workers, shown in the last row of
Table 8.2. Gender as a category, in this case, thus identifies heterogeneity
with respect to wage earnings in the population: women earn less than
men; hence, averaging across male and female workers is incorrect.
282 Econometrics far developing countries
Table 8.2 Distribution of weekly wage earnings (per cent by gender in each
income group)
Weekly wage earnings Gender
(lndian Rs) Male Fe mal e Total
(!) (2) (3) (4)
Up to 70 16.5 54.6 24.5
71-150 39.3 23.6 36.0
151-300 22.8 12.7 20.7
300 + 21.4 9.1 18.8
Total 100.0 100.0 100.0
Average weekly income 182.9 102.0 165.9

Recall that regression is the loci of conditional means or averages of


the dependent variable for given values of the explanatory variable(s).
The last row in Table 8.2 is, therefore, the sample regression of wage
income on gender as it gives the conditional average of wage income given
the gender category; that is:
Average (W G = Male) = 182.94
1

Average (W G = Female) = 101.99


1 (8.1)
where, W = weekly wage income; G = gender.
In Chapter 4, we discussed the regression of a measurement (numer-
ical) variable on another similar variable. Consequently, we could talk
about a regression line or a curve which depicts the average of the depen-
dent variable at each point in the continuous range of the explanatory
variable. If the regressor is a categorical variable, however, we can only
consider averages for each category: a series of distinct conditional aver-
ages. But the point to note is that we are essentially talking about the
same thing: the locus of conditional averages, either in the form of a
continuous line or a curve if the explanatory variable is numerical and
continuous, or in the form of a discrete series when the explanatory vari-
able is categorical.
The latter type of a discrete regression is not only applicable to cate-
gorical variables. A discrete numerical explanatory variable such as family
size will produce a similar result: we shall have average per capita income,
for example, at each discrete point of family size. But, in this case, we
can talk of increasing or decreasing conditional averages with increasing
values of the explanatory variable - a positive or a negative relationship.
By contrast, categories cannot always be ordered, for example, for the
gender variable (a nominal categorical variable). Here, we can only
analyse the differences in conditional averages to answer questions such
as whether, on average, men earn more than women. Obviously, when
our categorical variable can be ordered, for example, educational achieve-
ment, it is possible to speak of positive or negative association: income
rises with higher education achievement.
Categories, counts and measurements 283
How do we use categorical variables as regressors in our familiar linear
regression model? To do this, we make use of dummy variables, with
which you are familiar from Chapter 6. In the case of our example of the
relation between weekly wage income and gender we construct a numer-
ical variable, say D, so that:
D =O if G = male
D = 1 if G = female (8.2)
where D is a dummy variable denoting the categorical variable gender,
G. The difference between D and Gis that the latter names two distinct
categories while the former is numerical in nature. In other words, we
can add up D or multiply it with sorne other number, but we cannot do
the same with G. As far as G is concerned, we can only count the number
of observations in each category. In fact, in this case, adding up D yields
the same result as counting the number of women within the sample. We
can now write the sample regression as follows:
Average ( W 1 G) =average ( W 1 D )
= b1 + b2 D
which implies:
Average ( W 1 G = Male ) = Average ( W i D = O )
= b¡ + b2 o= b¡
Average ( W 1 G = Female ) = Average ( W D = 1 )
1

= b¡ + b2 1 = b¡ + b2
In other words, b 1 is the average wage income of the male workers, i.e.
b1 = 182.94, and b 2 is the difference between average female and male
wage incomes: b 2 = 101.99 - 182.94 = -80.95. This result may be proved
more formally (see exercise 8.8). Note that we are talking about sample
averages (or, more precisely, sample means). The regression model, there-
fore, can be written as follows:
w = 131 + 132 D + E (8.3)
where 13 1 and 13 2 are the population parameters of the model: respectively,
the mean earnings of male workers and the mean difference in earnings
between female and male workers.
However, not surprisingly, the distribution of weekly wage earnings of
the 261 workers is skewed, as can be seen from Table 8.2. The skew in
the distribution is clearly discernible - while the average weekly income
is about Rs 166 (given in the last row), more than 60 per cent of the
workers earned below Rs 150. The standard assumptions about E in the
classical regression model are obviously not applicable here. Therefore, a
transformation of the dependent variables is called far: once more, the
logarithmic transformation of W does the trick. Hence, we rewrite the
model (8.3) as:
ln(W) = 13 1 + 13 2 D + E (8.4)
284 Econometrics for developing countries
As explained earlier, in regression model (8.4) the constant term, 13 1, gives
us mean log earnings of male workers while the slope coefficient, 13 2 , states
the mean difference in log earnings between female and male workers.
This is typical of modelling categorical variables with dummy variables:
one of the categories is used as the benchmark represented by the constant
term, and the other categories are then compared (through differencing)
with this reference category by means of their corresponding slopes. Box
8.1 tells you why this is the best way to use dummy variables in regression

Box 8.1 Dummies and the constant term


In principle, there are two ways in which dummies can be made to
represent a categorical variable with a given number of mutually
exclusive categories. One way is to use a dummy variable for each
category. Hence, in the case of the gender variable, we could use a
dummy, DM, denoting men (DM = 1 if the person is male; O other-
wise) and another dummy, Dp, denoting women (Dp = 1 if the person
is female; O otherwise ). The other way is to use one fewer dummy
than the number of categories because, if all dummies equal O, the
remainder category which depicts the benchmark will be selected.
Which method is preferable? Clearly, the latter method saves us
from defining an additional, rather redundant, dummy variable. This
in itself is a good enough reason to use ( C - 1) dummies when there
are C categories. But there is a further reason to use ( C - 1) dummies
for C categories. This has to do with their use in regression analysis.
To see this, consider again the example of the gender variable.
Suppose we prefer to use two dummies, DM and Dp, and decide to
feature them both in a regression line along with a constant term. As
a result, the regressors (including the constant term) will now be
orthogonal. Indeed, you can easily verify that DMi + Dp; = 1, for all
i = 1, 2, 3, ... n, where n is the sample size. Now, since the constant
term is a variable X 1 such that X 1; = 1 for all i, it follows that a per-
fect linear relation exists between the dummies and the constant term:
X 1; - (DM; + Dp;) = O. Hence, if you want to use two dummies in
this case, you need to drop the constant term from the model.
It is generally less cumbersome to use (e - 1) dummies to depict
C categories as it avoids falling into the trap of rendering the regres-
sors orthogonal as a result of including the constant term in
the regression. The benchmark then depicts the category when all
( C - 1) dummies equal O. In the case of an ordinal variable it is
often useful to select the lowest (or highest) category as the bench-
mark.
Categories, counts and measurements 285
Table 8.3 Regression of log,(W) on D
Coefficient Standard errors t-statistic
D -0.703 0.119 -5.906
Intercept 4.941 0.055 90.383
Number of observations = 261; R 2 = 0.12

to denote categorical variables. Table 8.3 presents the results obtained by


estimating the log-linear model in equation (8.4).
Hence, b 1 , the estimate of 13 1, equals 4.942, the antilogarithm of which
equals 140 (rounded). As you know from Chapter 3, this is not an esti-
mate of the population mean of average weekly earnings of male workers,
but of its population median (and, incidentally, also of its geometric
mean). Table 8.2 shows that the mean of 183 (rounded) is well above this
median value, which shows that the data are indeed skewed to the right.
To obtain the median weekly earning of female workers we first have to
compute (b 1 + b2) = (4.94 - 0.70) = 4.24, the antilogarithm of which is
69.28, the estimate of the median weekly earnings of female workers. This
median is also significantly below the mean of 102, again showing that
these data are also skewed to the right.

Exercise 8.2
Using Table 8.3, formally test the hypothesis b2 = O. Does your conclusion
confirm that gender matters in terms of explaining earning differences?

The categorical variable for gender is a dichotomous variable, which


implies that there are only two categories: male and female. As a result,
a single dummy variable is sufficient to represent this categorical variable.
If our categorical variables have more than two categories (say, a total
number of C categories), we need (C - 1) dummies, as explained in Box
8.1. For example, education achievement of the workers in our data set
has four categories: below primary, primary, secondary, and post-
secondary education. A regression between log income (W) and education
(E), therefore, requires three dummy variables defined as follows:
D 2 = 1 if E = primary, D 2 = O otherwise
D 3 = 1 if E= secondary, D 3 = O otherwise
D 4 = 1 if E= post-secondary, D 4 = O otherwise. (8.5)
The matrix in Table 8.4 shows how the three dummy variables D 2 , D 3 , D 4
together represent the four categories of education. Each category of
education is represented by a unique combination of zeros and ones. In
the case of 'below primary' education all dummy variables are zero. Thus,
'below primary' education serves as the reference category, the bench-
mark, to which other educational categories are compared.
286 Econometrics for developing countries
Table 8.4 Categories of education
Educational achievement
Below primary o o o
Primary 1 o o
Secondary o 1 o
Above secondary o o 1

Table 8.5 Regression of log(W) on educational levels


Coefficient Standard error t-statistic
D2 0.081 0.134 0.604
D3 0.531 0.129 4.107
D4 0.827 0.204 4.059
Intercept 4.622 0.066 70.491
Number of observations = 261; R 2 = 0.10

The regression model of income on education is specified as:


ln W = ¡3 1 + ¡3 2 D 2 + ¡3 3 D 3 + ¡3 4 D 4 + E (8.6)
where, ¡3 1 = average log income of workers with below primary educa-
tion; ¡32 = difference of average log income of workers with primary
education from ¡3 1; ¡3 3 = difference of average log income of workers with
secondary education from ¡3 1; ¡3 4 = difference of average log income of
workers with post-secondary education from ¡3 1• Hence:
Mean ln(W) = ¡3 1 if E = below primary
= ¡3 1 + ¡32 if E = primary
= ¡3 1 + ¡3 3 if E = secondary
= ¡3 1 + ¡3 4 if E = post-secondary
The results of estimation of this model are given in Table 8.5.

Exercise 8.3
Using Table 8.5, formally test the hypotheses ¡3j = O, for j = 2, 3, 4. What
do you conclude in terms of the importance of educational level on weekly
earnings?

You should have found that there does not appear to be a significant
difference between average earnings of workers with primary and below
primary education, but both secondary and post-secondary education
clearly matter.
Categories, counts and measurements 287
Exercise 8.4
In exercise 8.1 you were requested to construct a categorical variable to
depict income per capita categories for different countries. Select life
expectancy as an additional (numerical) variable from the data set
SOCECON and:
1 compare mean- and order-based statistics of life expectancy for each
category of the income variable and construct the corresponding
comparative box plots;
2 construct dummy variables to represent this categorical income variable;
3 regress life expectancy as dependent variable on the dummies thus
constructed.
How does your analysis under (1) compare with your regression results
obtained in 3?

8.3 CONTINGENCY TABLES: ASSOCIATION BETWEEN


CATEGORICAL VARIABLES
We can now begin to put gender and education together to analyse their
joint influence on wage income. In other words, let us now consider a
multiple regression of a measurement variable such as income on the
dummy variable representations of the categorical variables gender and
education. But befare we do so we need to do sorne more groundwork
and develop ways to examine the association between two categorical vari-
ables. As an example, we ask whether educational achievement is itself
associated with gender.

Contingency tables
First, we take up the issue of association between gender and education
which, if it exists, will play a role in the partial associations with income
(recall Chapter 5). To do this we use a contingency table. A two-way con-
tingency table presents a joint frequency distribution of two categorical
variables. Table 8.6 is the contingency table of gender and education,
listing frequencies (or counts) of 261 workers jointly by gender and by edu-
cation. A contingency table such as is given in Table 8.6 is used to examine
whether or not two categorical variables are stochastically (statistically)
independent (see Box 8.2). We need a table of cross-tabulations because,
with categorical variables, we cannot compute numerical summaries like a
covariance ora coefficient of correlation straight away. But we can use the
frequencies to compute fractions of counts or proportions which, as we shall
see, serve as estimates of the corresponding probabilities.
To say that two categorical variables are stochastically dependent does
not imply any statement about causation. It merely states that they
are associated with one another. Hence, when investigating stochastic
288 Econometrics far developing countries

Box 8.2 Stochastic independence


Two events A and B are stochastically independent if
P(A j B) = P(A) or P(B j A) = P(B)
In other words, A and B are stochastically independent if occurrence
of one of the events does not change the probability of occurrence
of the other event.
Thus, gender and education of a worker are stochastically indepen-
dent if the probability of a certain level of education achievements
of a randomly selected worker (event A) remains the same irre-
spective of whether the worker is male or female (event B). Or,
equivalently, the probability that a randomly selected worker is male
or female is the same irrespective of the educational category from
which the worker is selected.
The definition of conditional probability implies:
P(A j B) = P(A and B )/P(B)
or
P(A and B) = P(A j B) · P(B) { = P(B j A) · P(A) }
Therefore, if A and B are independent: P(A B) 1 = P(A), it follows
that:
P(A and B) = P(A)·P(B)
In other words, if gender and education are independent then:
P(G =g and E= e) = P(G = g)·P(E =e)
for all combinations of g and e, where G = Gender (g = male
or
female); E= education (e= below primary, primary, secondary, or
post-secondary ).

Table 8.6 Contingency table of gender education


Education
Gender Below primary Primary Secondary Above secondary Total
Mal e 111 40 44 11 206
(42.5) (15.3) (16.9) (4.2) (78.9)
Female 36 6 7 6 55
(13.8) (2.3) (2.7) (2.3) (21.1)
Total 147 46 51 17 261
(56.3) (17.6) (19.5) (6.5) (100.0)
Note: Figures in parentheses indica te cell counts as per cent of total counts ( = 261 ).
Categories, counts and measurements 289
dependence between two (or more) categorical variables we do not neces-
sarily imply any direction of causation. If we have good reason to believe
that one variable depends causally on the other, it is preferable to model
the data likewise by making explicit that one variable depends on the
other(s). This we shall do in the context of logit modelling in the next
chapter. Here, our interest is to see whether two categorical variables
'covary' in the sense that they are not statistically independent.
The test of independence that we develop is based on a comparison of
the observed counts or frequencies of the contingency table with what
would be the expected frequencies if the two variables were independent.
To work out what are the expected frequencies under the hypothesis
of independence requires that we make use of the property that P(A and
B) = P(A) ·P(B) if A and B are stochastically independent in the context
of the binomial distribution.

The binomial distribution


The binomial probability distribution is concerned with modelling a
dichotomous variable whose outcome in a single trial can be one of two
possibilities only. The common analogy used for discussing the binomial
distribution is that of tossing a coin which results in either a 'head' or a
'tail'. A practica! example is our sample worker being male or female.
Let p be the probability that a randomly selected worker is female and,
hence, (1 - p) is the probability that a randomly selected worker is male.
The binomial distribution model gives us the probabilities of the number
of women, F, in an independent random sample of size n. The value of
F can be anywhere between zero (no women in the sample), and n (ali
sample workers are women). Hence, F can assume any of the values
(O, 1, 2, ... n). The variable F is called a binomial variable, and the prob-
ability p is called the binomial probability.
To express the probability that F = f (in a sample of size n) we proceed
as follows. The probabilities of sampling f female workers and (n - f) male
workers in a particular sample are, respectively, pf and (1 - p )n-f, and, hence,
pf(l - p )n-f is the probability associated with this sample. But there are
severa] ways of getting this type of sample. For example, in a sample of size
three (i.e. n = 3), one can get two women (F = 2) and one man in three
different ways: (male, female, female ), (female, male, female) or (female,
female, male ). Each of these samples has a probability of p 2 (1 - p )3- 2.
Thus, the probability of F = 2 in a sample of 3 equals 3 p 2 (1 - p )3- 2 . In
general, the number of ways one can get f women and (n - f) men in a
sample is given by the following convenient algebraic formula:
Cp = number of ways f women and (n - f) men can
be obtained in a sample of size n
= n!ll f! (n - f)! ) where k! = k(k - l)(k - 2) ... 2.1;
290 Econometrics far developing countries
and, hence, the probability distribution of F can be written as:
P (F = f) = e¡ pf (l - p yn-t f = 0,1,2, ... , n (8.7)
which is the well-known binomial distribution, denoted by B(n,p), where
n, the sample size, and p, the binomial probability, are the parameters of
the distribution. Thus, if F has a binomial distribution B(n,p ), then its
probabilities for F = 0,1,2, ... , n are given by the expression above. The
theoretical mean and variance of a binomial distribution is given as
follows:
F ~ B(n,p) (8.8)
E(F) = n p; V(F) = n p (l-p)
Hence, the mean of a binomial distribution with parameters n and p is
the product of the two parameters: its sample size multiplied by the bino-
mial probability. Now, it can be shown that the sample proportion is an
unbiased and mínimum variance estimator of the probability p, a prop-
erty we can use to work out the expected frequencies under the hypothesis
of independence.

Contingency chi-square test of independence


Let us return to our contingency table (Table 8.6). Each cell in Table 8.6
corresponds to a specific gender and education category. Let us consider
a particular cell - male and below primary. Hence, only if a worker is
male and has below primary education will he be counted in this cell;
otherwise, not. This is obviously a dichotomous situation. Now, recall that
the sample of workers is independently and randomly selected. Therefore,
we have 261 randomly selected workers each of whom either falls into
this specific cell or <loes not. The situation is similar to that of a binomial
distribution discussed above: that is, the cell frequency can be modelled
by a binomial distribution. Hence, we can treat the cell frequency as
a binomial variable with binomial probability equal to P( G = male and
E = below primary). The observed frequency (OBS) in the cell is 111;
its expected frequency (EXP) will, according to equation (8.8) be equal
to n P(G = male and E = below primary). Now, under the hypothesis
of independence, we get P(G = male and E= below primary) = P(G =
male)·P(E = below primary) and, hence:
EXP(G = male and E= below primary) =' n P(G = male and
E = below primary)
= 261 P(G = male)
·P(E = below primary)

P( G = male) and P(E = below primary) are also binomial probabilities


corresponding to binomial variables: respectively, the number of male
Categories, counts and measurements 291
Table 8.7 Computation of contingency chi-square statistic
Observed Estimated Expected cell Relative squared
cell frequency probability frequency deviations
(O) (P) (261.P) (O - E) 2/E
111 (206/261)-(147/261) 116.02 0.218
40 (2061261H461261) 36.31 0.378
44 (206/261)'(51/261) 40.25 0.349
11 (206/261)'(17 /261) 13.42 0.436
36 (55/261)'(147/261) 30.98 0.814
6 (55/261)'(46/261) 9.69 1.407
7 (55/261)'(51/261) 10.75 1.307
6 (55/261)'(17/261) 3.58 1.632
261 6.538

workers and the number of 'below primary' workers. Since the sample
proportions are good estimators of the corresponding binomial probabil-
ities, we can use the corresponding sample proportions of male workers
(206/261) and 'below primary' workers (147/261) as estimators of these
probabilities. Hence, the estimated expected cell frequency for ( G = male
and E= below primary), under the hypothesis of stochastic independence,
is equal to 261 (206/261)-(147/261) = 116.02, which compares with an
observed frequency of 111. The same steps can be repeated for the rest
of the cells of the contingency table to obtain the estimated expected
frequencies under the hypothesis of independence. The third column in
Table 8.7 lists all estimated expected cell frequencies.
If the hypothesis of independence is correct, then these expected
frequencies should be very clase to the actually observed frequencies in
the sample, i.e. the first and the third column of Table 8.7 will be very
clase to each other. The more the third column deviates from the first
column, the less likely it is that both variables are indeed stochastically
independent. What we need, therefore, is a numerical tool to summarise
the deviations of the expected frequencies (EXP) from the observed
frequencies (O BS). This is done as follows:
2
x2 = 2: [(OBSE~:XP) ~ ( 8.9)

where the summation is over all the cells in the contingency table.
The resulting measure is called the contingency chi-square statistic. It
is easy to see that the more the expected frequencies deviate from the
observed frequencies, the larger will be the value of the statistic since it
is based on the squared differences between the two sets of frequencies.
The larger the value of the statistic, therefore, the more the hypothesis
of independence is suspected. But how large a value of this statistic is
large enough to reject the hypothesis of independence? Such a judgement
is based on how unlikely (in the probability sense) is the large value of
292 Econometrics for developing countries
the statistic. Now, the sampling distribution of this statistic is a chi-square
distribution with degrees of freedom equal to d = (e - l)(r - 1), where e
is the number of columns and r the number of rows in the contingency
table. This explains the name of the summary measure in equation (8.9).
So all we need next is to find the cut-off points (critica! values) in the cor-
responding chi-square distribution at which there is only 5 per cent or
1 per cent probability (level of significance) for the statistic to have a larger
value. If the computed value of the statistic is larger than the critica! value,
then we reject the hypothesis of independence at the corresponding level
of significance. Note that computer software often provides the upper-tail
probability of the computed value, i.e. the probability of the statistic being
larger than the computed value. In that case, we reject the hypothesis of
independence if the upper-tail probability is less than 5 per cent or 1 per
cent. This procedure to test the hypothesis of independence of two cate-
gorical variables is called the contingency chi-square test of independence.
In our present example, the value of the test statistic computed at
the bottom of the fourth column is 6.54. The degrees of freedom are
(4 - 1)·(2 - 1) = 3. The critica! value of a chi-square distribution with
three degrees of freedom is 7.81 at 5 per cent level of significance. Since
the computed test statistic of 6.54 is below the critica! value we cannot
reject the hypothesis of independence between gender and education of
workers at the 5 per cent level of significance.
The test of independence described above checks for the existence of
association. The larger the value of the test statistic, the greater is the
evidence of association. The same chi-square statistic can also be used to
develop a measure of association. One such measure is given by Cramer's
V, defined as follows:
2
V= J(OBS- EXP) . 1
(8.10)
EXP n·(k-1)

where n is the total frequency and k is the mínimum of the number of


rows and columns. The value of V líes between O and 1, both inclusive.
In our case, Cramer's V measure of association between gender and educa-
tion is equal to [6.5375/1261.(2 - l)JF' = 0.16, which indicates that the
association is rather weak.

Exercise 8.5
Using exercises 8.1 and 8.4 and the data set SOCECON (world socio-
economic data), define a categorical variable which picks out the countries
of Sub-Saharan Africa from among the developing countries. Investigate
whether this variable and the categorical income variable are statistically
independent or not, using both the contingency chi-square test of inde-
pendence and Cramer's V statistic.
Categories, counts and measurements 293
8.4 PARTIAL ASSOCIATION AND INTERACTION
In our example we found that both gender and education are related to
wage income. To distinguish between their separate effects on wage income,
we shall make use of the familiar concept of partial association. But we also
need another concept which, initially, is a little difficult to understand but
matters a great deal when working with categorical variables. This is the
concept of interaction between two (or more) categorical variables in their
effects on a measurement variable: for example, the differences in partía!
associations of income with education across the gender categories.
Let us start with the question of the partial associations of income with
gender and education. For convenience of exposition, we collapse the four
categories of education into two categories - 'below secondary', and 'sec-
ondary and above'. The partial association between weekly wage income
and education is the association between these two variables, keeping the
third variable, gender, fixed at a specific category. For example, consider
the contingency table of income groups and educational categories for
male workers in Table 8.8(a). The degree of association in this contingency
table is the association between income and education for male workers.
Therefore, it is a partial association between income and education.
Similarly, the degree of association between income and education in
the contingency table for female workers, Table 8.8(b ), is also a partía!
association. The chi-square statistics indicate that the hypothesis of inde-
pendence between income and education can be rejected at the 1 per cent

Table 8.8 Partial associations of weekly wage income and education


Weekly wage income Below To or above Total
secondary secondary
education education
(a) Male workers
Up to 70 29 5 34
71-150 68 13 81
151-300 30 17 47
Over 300 24 20 44
Total 151 55 206
Chi-square (3 df) = 17.26
Upper tail probability = 0.001
Cramer's V = 0.29
(b) Female workers
Up to 70 28 2 30
71-150 7 6 13
151-300 5 2 7
Over 300 2 3 5
Total 42 13 55
Chi-square (3 df) = 12.20
Upper tail probability = 0.007
Cramer's V = 0.47
294 Econometrics far developing countries
level of significance for both male and female workers. In other words,
there is evidence of association in both cases, which is not at all surprising.
Higher education is likely to bring better-paid jobs. What is more inter-
esting is that the sample association between income and education is
comparatively stronger among female workers than among male workers
as indicated by Cramer's V measures of association. In other words, the
partial association between income and education varies with gender.
Hence, the way in which education relates to income depends on whether
we are dealing with men or women: we say that education and gender
interact with each other in their respective association with income.
Take a look at Table 8.9, which presents the average log income of
workers by education and gender. The entries in the table are the condi-
tional averages of the logarithm of weekly earnings for given gender and
education categories. The rows in the table are the partial regressions of
log income on education, because each row presents average log income
by education, keeping gender fixed. Similarly, the columns present the
partial regressions of log income on gender, keeping education fixed. The
interaction effect shows up in the way the partial regressions differ.
For example, male workers with secondary or above education fetch 0.5
( = 5.31 - 4.81) higher average log income as compared with other male
workers with 'below secondary' education, while for female workers the
similar difference is 0.84 ( = 4.88 - 4.04). In other words, the effect of
higher education is different between males and females. The interaction
effect can also be seen the other way, that is, how the average log income
differs between gender groups within each educational category. The
differences are 0.77 ( = 4.81 - 4.04) and 0.43 ( = 5.31 - 4.88) for lower
and higher education categories, respectively; that is, the gender differ-
ences in average log income is less among workers with higher education.
In the absence of interaction effects these differences in average log
income, between gender or between education, would be the same.

Exercise 8.6
Using the data set SOCECON, investigate the partial associations and the
interaction effects of the categorical income variable and the variable
denoting Sub-Saharan African countries on life expectancy. In each case,
explain carefully what each of these concepts measures in this concrete
example.

Table 8.9 Average log wage income by gender and by education


Below secondary education To or above secondary education
Mal e 4.81 5.31
Fe mal e 4.04 4.88
Categories, counts and measurements 295
8.5 MULTIPLE REGRESSION ON CATEGORICAL VARIABLES
Our next step is to develop an appropriate multiple regression model in
arder to study the relationship of weekly wage income with gender and
education such that we capture both the separate influence of each of
these explanatory variables net of the other and their interaction effect
on income. To be able to use the regression model we need to construct
dummy variables since our explanatory variables are categorical in nature.
Let DG and DE be the dummy variables for gender and education.
DG = O if gender = male
= 1 if gender = female (8.11)
DE = O if education < secondary
= 1 if education ~ secondary (8.12)
DGE = DG·DE
= 1 if gender = female and education ~ secondary
= O otherwise (8.13)
where DGE is the interaction dummy between gender and education.
The multiple regression model can now be specified as follows:
/'--.._
ln(W) = 13 1 + ¡3 2DG + ¡3 3DE + ¡3 4DGE (8.14)
where, ¡3 1 = average log income for men with 'below secondary' educa-
tion which is the benchmark category; 13 1 + ¡3 2 = average log income for
women with 'below secondary' education; 13 1 + ¡3 3 = average log income
for men with 'above-secondary' education; 13 1 + 13 2 + 13 3 + 13 4 = average
log income for women with 'above-secondary education.
We could now proceed by estimating this model with least squares
regression. However, befare we do this, it is instructive to show how the
regression can be deduced from Table 8.9 by taking row and column differ-
ences. This is done in Table 8.10.
The row and column differences in Table 8.10 clearly show how the
interaction effect comes into play. Note, in particular, that the difference
of the differences (right-hand bottom comer) gives us the interaction

Table 8.10 Breakdown of average log income into components


Below At or above Difference
secondary secondary
education education
Male 4.81 5.31 = 0.50 (= [33)
( = í31)
Female 4.04 4.88 = 0.84 = 0.50 + 0.34
(= [33 + [34)
Difference = ---0.77 = -0.43 = -0.34 (= [34)
(= í32) = -0.77 + 0.34
(= í32 + [34)
296 Econometrics for developing countries
Table 8.11 Regressing log income on gender and education
Coefficient Standard error t-statistic
DG -0.77 0.13 -5.92
DE 0.50 0.12 4.63
DGE 0.34 0.26 1.29
Intercept 4.81 0.06 79.37
Number of observations = 261; R 2 = 0.21

effect directly. Therefore, if there is no interaction, column and row


differences add up to zero. In this case, there is sorne interaction. To see
whether it is significantly different from zero or not, it is best to turn to
the estimation of model (8.14) with least squares regression. This is done
in Table 8.11.
The regression results reproduce the results obtained in Table 8.10. Take
sorne time to verify this by carefully comparing these regressions with the
column and row margins (depicting differences) of Table 8.10. The regres-
sion model, therefore, does not add much except for the fact that you can
carry out the tests of significance more easily. Note that the interaction
dummy turns out not to be significantly different from zero at the 5 per
cent significance level.
As we have already seen in Table 8.10, the regression coefficients can
be easily read out from Table 8.9. This is only possible, however, when
the explanatory variables are all categorical.
What makes the regression modelling by means of dummy variables
particularly useful is that it allows us to combine measurement variables
along with categorical variables in the same model. For example, up to
now we have not yet considered the age of the workers in analysing their
weekly earnings. But age obviously matters. To the extent that workers
earn more with age as they acquire skills or experiences, variations in the
age of the workers must be at least part of the reason for the variations
in the earnings. Let us, therefore, incorporate the age as a variable in our
model:
(8.15)
The estimated regression is given in Table 8.12. This model fits quite well.
The residuals behave approximately normal: the skewness-kurtosis test
turns out to be insignificant at the 5 per cent level (results not presented
here). By all indications, the regression model in (8.15), therefore, appears
to be a satisfactory specification.
Interestingly, the interaction coefficient turns out to be significantly
different from zero at the 5 per cent level and, moreover, it takes on a
higher value of 0.52 as compared to 0.34 befare we included age into the
model! All other coefficients are also significant at the 5 per cent level.
Obviously, omitting the age factor blurred the interaction effect. What we
Categories, counts and measurements 297
Table 8.12 Regression of log income on gender, education and age
Coefficient Standard error t-statistic
DG -0.75 0.11 -6.90
DE 0.50 0.10 5.06
DGE 0.52 0.22 2.33
AGE 0.03 0.01 10.38
Intercept 3.72 0.11 31.93
Number of observations = 261; R 2
= 0.45

have achieved in the final regression analysis in this section is first to


remove the impact of age variation among the workers on the variation
in weekly wage earnings, and then to look deeper into the variation by
identifying heterogeneity among the workers with respect to gender and
income. The R 2 value indicates that about 45 per cent of the variation
in weekly wage earnings is explained by age, education and gender.
Significantly, we found that there was no evidence of association between
gender and education (recall the result of the contingency chi-square test).
This is probably not so much indicative of women having the same access
to education as men, but rather that men also have limited access to educa-
tion because of the socioeconomic handicap suffered by working-class
households which, in turn, reduces the gender gap in education. The nega-
tive coefficient of the gender dummy shows that female workers earn less
than male workers irrespective of education. But a redeeming feature is
the positive coefficient of the dummy for the interaction between gender
and education: the incremental earnings of female workers with higher
education is higher than that of the male workers. In other words, the
earning benefits of higher education is higher for women. But this only
means that the gender gap in earnings reduces with higher education: on
average, women still earn less.
The introduction of the age variable changed the picture quite dramat-
ically with regard to the interaction dummy between gender and educa-
tion. This illustrates the argument we made in Chapter 6 that we should
beware of omitting an important relevant variable from the model since
our coefficient estimates will end up being biased as well as less precise.
In fact, this example illustrates how the introduction of an important rele-
vant variable increases the precision of the estimates of the variables
already included in the model. Note how introduction of the age variable
improved the t-statistics of all dummy variables included in the model.
Compare Table 8.12 with Table 8.11. Furthermore, we also note that the
bias produced by omitting the age variable mainly affected the interac-
tion variable DGE, the coefficient of which increases markedly. The other
slope coefficients, however, hardly changed at ali.
Obviously, we should have known all along that age matters. In fact,
in Chapter 7 we discussed the regression of the logarithm of wage income
298 Econometrics far developing countries
on the age of workers. But the structure of the argument in this chapter
reflects pedagogical concerns, not research priorities. Hence, we started
step by step, introducing our categorical variables one by one, moving on
to multiple regression with two dummy variables, and ending up with a
regression which features dummies along with a measurement variable as
regressors in the model. In actual research we would have started with a
general model which included the age variable from the beginning along
with a string of dummy variables and subsequently we would have checked
whether any of the variables could be dropped from the equation.
What about the interaction between a categorical variable and a quan-
titative variable? Is this possible? Yes, it is and, in fact, we already dealt
with such interaction in Chapter 6. Note, for example, that our specifica-
tion (8.15) only includes intercept dummies, although one of them is an
interaction term. The interaction between a categorical variable and a
quantitative variable is done with the slope dummy. In general, the tech-
nique consists in introducing relevant product terms between the dummy
variable and the quantitative variable in question. For example, a model
of multiple regression of Y on X and D (respectively, a measurement vari-
able and a dummy variable) along with the interaction effect of X and D
can be specified as follows:
(8.16)
where ¡3 4 captures the interaction effect. There is no need to give exam-
ples here since Chapter 6 showed you how to do this. What is useful,
however, is to go back to Chapter 6 and see how each of the dummies
used reflect categorical variables (policy regimes; time periods; heteroge-
neous sub-samples ).

Exercise 8. 7
Show how regression analysis can be used to investigate the relation
between life expectancy as dependent variable and the categorical income
per capita variable and the Sub-Saharan Africa identification variable as
explanatory variables. Check whether your results conform with those
obtained in exercise 8.6.

8.6 SUMMARY OF MAIN POINTS


1 Categorical variables allow us to classify our data with respect to
sorne qualitative criterion in a set of mutually exclusive categories. A
qualitative variable is dichotomous if it contains only two mutually
exclusive categories. We cannot measure a categorical variable, but we
can count how many observations fall into a particular category. This
allows us to perform quantitative calculations with respect to frequen-
cies of qualitative variables.
Categories, counts and measurements 299
2 To include qualitative variables into regression analysis as explanatory
variables we make use of dummies. In general, we use (C-1) dichoto-
mous dummies to represent a categorical variable with C categories.
Each dummy takes on the value 1 if the observation belongs to its
category, and zero otherwise. The category without a corresponding
dummy variable is the benchmark or reference category which is
selected when all dummies equal zero.
3 A contingency table allows us to study the association between two
(or more) categorical variables. This is done by verifying whether the
variables are statistically independent from each other. The test
consists in comparing observed with expected counts (i.e. frequencies),
where the expected counts are derived with the aid of the binomial
distribution, assuming randomly sampled observations. The relevant
test is the contingency chi-square test of independence. A derived
measure of association is Cramer's V which líes between zero and one.
4 When looking at the relation between a quantitative dependent vari-
able Y and two (or more) categorical variables, W and Z, we take
account of partial associations and interaction effects. The partial asso-
ciation between Y and W is the association between them, keeping Z
fixed within a given category. If such partial associations between Y
and W varíes for different categories of Z, we say that there is inter-
action between W and Z in their effects on Y.
5 The use of dummy variables associated with two or more categorical
variables allows us to study partial association and interaction effects
in the context of multiple regression. Interaction dummies are obtained
by multiplying dummies corresponding to different categorical vari-
ables. This procedure allows us to test formally whether interaction is
present or not using the familiar t-test (or an F-test if we want to drop
more than one interaction term from the regression).
6 A slope dummy is nothing else but an interaction variable between a
dummy variable and a quantitative variable. Its slope measures the
interaction effect and its t-statistic shows whether it is significant
or not.

ADDITIONAL EXERCISES

Exercise 8.8
Consider the regression model
y = ¡3¡ + ¡32 D + E
where Y is income and D a dummy variable for gender (D = 1 for female
and D = O for male). Use the formulae for the least squares estimators
to show that
300 Econometrics for developing countries
- - -
b 1 = Y m and b 2 = Y 1 - Y m
where the m and f subscripts denote male and female respectively. (The
solution to this exercise is given as Appendix 8.1.)

Exercise 8.9
Repeat your analysis of life expectancy carried out in this chapter but
now including dummy variables to separately identify Sub-Saharan Africa,
Asia, North Africa and the Middle East, Eastern Europe and developed
countries. Compare your results with those obtained earlier.

Exercise 8.10
Using the data in the PRODFUN file, calculate labour productivity and
classify firms as having low, medium and high productivity. Is the level of
labour productivity independent from the sector in which a firm operates?

APPENDIX 8.1: SOLUTION TO EXERCISE 8.8


The model is
(A.8.1)
We know that
2,(D - D)Y
b2 = 2,(D - D)2
(A.8.2)
Now

:. D - D = nm if D = 1
n

-_ - --L
n, if D =O
n (A.8.3)
where n1 and nm are the number of women and men respectively. The
denominator in equation (A.8.2) is therefore given by:

(~)
2
-n
_
m n
+n1 (nm)
n
2
-_ !!d!J
n
(A.8.4)
Categories, counts and measurements 301
The numerator is:
n nm ~
2: (D -
i=l
D)Y; = 2: (D; -
i=l
D)Y; + 2.J (D; - D)Y;
i=l

(A.8.5)

Substituting equations (A.8.4) and (A.8.5) into (A.8.2):

(A.8.6)

the intercept is given by:


b1 =Y- b2 D

=Y- nmn [Y - Y] f m

(A.8.7)
9 Logit transformation, modelling
and regression

9.1 INTRODUCTION
Up to now, the dependent variables in the economic models we discussed
were all measurement variables. But what if the dependent variable we
are interested in is categorical in nature? For example, we may be inter-
ested in investigating the main determinants of home ownership in an
urban setting. Alternatively, our interest may be to find out why sorne
rural labourers succeed in obtaining permanent jobs while others have to
depend on temporary work or casual jobs. In both examples the depen-
dent variables are categorical in nature. In fact, they are both dichotomous
variables. This chapter looks at dependent categorical variables which are
dichotomous in nature and, hence, can be represented by dummy vari-
ables. Our focus, therefore, is on regressions in which the dependent
variable is a dummy variable.
When dealing with a dichotomous dependent variable our main interest
is to assess the probability that one or another characteristic is present.
Does a labourer have a permanent job, or not? Does the household own
its home, or not? What determines the probability that the answer is yes,
orno? It is the latter question which we try to address when dealing with
logit regression. This is what makes logit regression, despite many simi-
larities, essentially different from the linear regression models we have
discussed so far. In multiple regression, for example, we try to predict the
average value of Y for given values of the independent variables with the
use of a regression line. In logit regression, however, our interest is to
predict the probability that a particular characteristic is present. Hence,
we do not predict whether Y equals 1 or O; what we predict is the prob-
ability that Y = 1 given the values of the independent variables.
But what are logits anyway? We start this chapter by showing that
logits provide a convenient means to transform data consisting of counts
of a dichotomous variable. Hence we can show the connection between
proportions (or percentages), odds and logits as alternative summaries
of observed counts of a dichotomous variable. We argue that the logit
transformation is a handy and user-friendly tool for data analysis,
Logit transformation, modelling, regression 303
independently from its further usefulness in the context of logit model-
ling and regression.
Next we turn to logit modelling in the context of multiway contingency
tables. You have already come across the concepts of contingency tables
and dependence or independence of categorical variables in the previous
chapter. In this chapter, we show how this analysis can be extended to
investigate the effects of one or more categorical variables (the indepen-
dent variables) on a dichotomous response variable (the dependent
variable). From here it is a small step to reach logit regression as a flex-
ible model to deal with regressions featuring a dummy variable as
dependent variable, which is done in the closing sections of this chapter.
The final section summarises the main points.

9.2 THE LOGIT TRANSFORMATION


When dealing with a dichotomous categorical variable we count the
frequency of a given characteristic in relation to the total number of obser-
vations. Table 9.1, far example, depicts the prevalence of high profit
enterprises among informal sector enterprises in Tanzania by sector,
comparing rural enterprises with those in Dar es Salaam, in 1991.
We can now calculate proportions, odds and logits relating to our dichoto-
mous categorical variable. In general, if N is the total number of observa-
tions and n is the number of observations which have the required property
(e.g. high profit enterprises), the proportion,p, is obtained as follows:
n
p=-
N (9.1)
For example, the proportion of high profit enterprises in the total number
of urban small scale manufacturing enterprises is given by

Table 9.1 High profit enterprises in Tanzania's informal sector, 1991


Number of high profit Total number of
enterprises enterprises
Rural Dar es Salaam Rural Dar es Salaam
Agriculture and fishing 4,482 1,116 77,634 11,810
Mining and quarrying 123 o 3,952 o
Manufacture 2,947 2,326 354,529 31,456
Construction 234 684 87,598 10,762
Trade, restaurants and hotels 11,723 10,368 525,730 142,041
Transport 1,143 682 41,508 2,324
C&P services 425 1,182 63,185 12,759
Source: Tanzania: The Informal Sector 1991, Planníng Commissíon and Minístry of
Labour, Table ENT25.
Note: Hígh profit enterprises are enterpríses wíth an annual profit of more than
Tsh 500,000.
304 Econometrics for developing countries
2326
p man,Dar = 31456 = 0.0739

or 7.39 per cent of all enterprises. Under conditions of random sampling,


a calculated proportion gives us an estimate of the probability of encoun-
tering a high profit enterprise in urban informal manufacturing.
The odds, O, is the ratio of the number of observations with the required
characteristic to the number of observations which do not have this char-
acteristic:

0=-n- (9.2)
n-N

p
(9.3)
(1 - p)

Far example, the odds favouring high profit enterprises in the urban small-
scale manufacturing sector in Tanzania are:
0.0739
O man,Dar = 1 - 0.0739 = 0.0798

In this case, the odds are very similar in value to the corresponding propor-
tion : 0.0798 as against 0.0739. The reason is easy to understand once we
take another look at equation (9.3). For small values of p, the denomi-
nator (1 - p) will be approximately 1 and, hence, the odds are approx-
imately equal to the corresponding proportion.
A logit, L, is obtained by taking the logarithm of the odds. Hence:
L = log (O) (9.4)
= log p/(1 - p) (9.5)
= log p - log (1 - p) (9.6)
In our example, the logit of high profit enterprises in urban informal manu-
facturing equals:
Lman,Dar = ln 0.0798
= -2.528
The difficulty sorne students have with logits is not to do with the way
it is calculated but with the question as to why we should bother computing
the logarithm of the odds in the first place. But, as we intend to show,
the logit transformation is a convenient way to deal with the analysis of
dichotomous categorical variables. Why?
Many people prefer to work with proportions (or, better still, percent-
ages). However, the problem with proportions is that they have clear
boundaries - a fioor of zero and a ceiling of one - which can cause trouble
when working with either very small or very large proportions. In the
Logit transformation, modelling, regression 305
0.3 tran

0.25

0.2

0.15

0.1

0.05

o
Rural Urban
s1a1a'M

Figure 9.1 Estimated probabilities (proportions)

case of our Tanzanian data, for example, most proportions of high profit
enterprises in the total number of enterprises turn out to be very small,
particularly in rural-based sectors. This can lead us astray when comparing
the variations in these proportions between urban and rural enterprises,
as shown in Figure 9.1. The comparative box plots show the variation
across sectors in the proportion of high profit enterprises in rural and
urban enterprises respectively. With the exception of urban transport as
a far-outlier, this figure conveys the impression that the variation across
sectors is fairly similar for rural and for urban enterprises. But, in fact,
the ratio of the highest to the smallest proportions in rural-based sectors
is far in excess of the same ratio for urban-based sector, even if we include
the outlier. You can easily verify that this ratio equals 4.6 for urban-based
sectors as against 21.6 for rural-based sectors. The comparative box plots
shown in Figure 9.1 effectively hide this greater internal variability among
rural sectors since all its proportions are small so that the box plot is
squashed against the fioor.
By contrast, Figure 9.2 shows the comparative box plots of the logits.
This plot shows much greater internal variability among rural-based
sectors as against urban-based sectors. Why is this? The answer lies in the
effects of the logit transformation on proportions. A proportion can range
from O to 1, that is:
306 Econometrics for developing countries
o

-1 tran

-2

-3

-4

-5

Rural Urban
srata™
Figure 9.2 Comparing logits

(9.7)
but the logit transformation stretches the tails of this distribution, and,
hence:
-oo~L~+oo (9.8)
where L = O (i.e. O = 1) corresponds to a proportion equal to 0.50. Hence,
logits bring out significant variation among small or large proportions far
better than can be seen from merely looking at the proportions them-
selves. Indeed, in our example, the box plot of the logits tells us that, with
the exception of urban transport, urban-based sectors are far more homo-
geneous with respect to the prevalence of high profit enterprise than
rural-based sectors.
The logit transformation, therefore, helps us to detect patterns in counts
or proportions of dichotomous categorical variables, particularly in cases
where the characteristic we are interested in is either very rare or highly
prevalent. Fortunately, in the middle range (corresponding to proportions
around 0.50), proportions and logits tell a similar story. Consequently, as
Tukey (1977: 509) puts it, logits are 'first-aid bandages' when working with
counted fractions since, on balance, they allow us to bring out patterns
in the data which may be obscured when using proportions or percent-
ages. For this reason, logits are useful exploratory choices for data analysis
with counted fractions of categorical variables. But, as we shall see below,
Logit transformation, modelling, regression 307
logits also improve our capacity to model data when the dependent vari-
able is a dichotomous categorical variable.

Exercise 9.1
Use Table 9.2 to compute the proportion, odds and logits of informal
sector enterprises which had existed less than five years by 1991 in both
rural and urban areas in Tanzania. Make comparative box plots for each
of these cases.

9.3 LOGIT MODELLING WITH CONTINGENCY TABLES


In economic analysis of development issues we prefer statistical models
that clearly specify which variable depends on others. Logit modelling
allows us to do this in the context of a qualitative dependent variable. In
this section we shall show how this can be done in the framework of
multiway contingency tables when all variables are categorical in nature.
You have already come across contingency tables in Chapter 8. In essence,
a contingency table is a cross-tabulation of counts across two or more
categorical variables. In the previous chapter we showed how these cross-
tabulations can be used to assess the degree of association between
categorical variables. We did not make any assumptions, however, about
the direction of causation between these variables. This section shows how
this can be done explicitly using logit modelling (Demaris, 1992).
Table 9.2 shows the cross-tabulation of counts of Tanzanian informal
sector enterprises by urban and rural location, by sector and by the age
of the enterprise for the year 1991. The counts for the age of the enter-
prises are grouped into enterprises less than five years old and those which
have existed for five years or more. The reason for grouping the age vari-
able with respect to a cut-off point of five years is that in Tanzania
structural adjustment policies were initiated during 1986 and hence, by
1991, had been operative for about five years. Consequently, the cate-

Table 9.2 Tanzanian informal sector enterprises by sector, location and age
Urban Rural
Less than 5 or more Less than 5 or more
5 years years 5 years years
Agriculture and fishing 39,788 24,687 34,471 43,163
Mining and quarrying 6,070 7,117 1,802 2,150
Manufacture 51,459 33,448 137,642 212,883
Construction 11,399 17,451 38,613 48,526
Trade, restaurants and hotels 308,312 96,881 341,017 179,995
Transport 5,411 2,460 31,674 9,181
C&P services 20,087 19,648 23,857 38,031
Source: Tanzania: The Informal Sector 1991, Planning Commission and Ministry of
Labour and Youth Development, ENT 5 (adapted).
308 Econometrics far developing countries
gorical variable of the age of the enterprise (more or less than five years
old) divides up the total number of enterprises by the type of policy regime
in operation when they were started: roughly, befare and after structural
adjustment policies were initiated.
Clearly, the data in this table can be used to explore questions about
structural changes in the spread of informal sector activities across sectors
befare and after the implementation of structural adjustment policies
and across the rural/urban divide. Note, however, that the table cannot
tell you much about the growth in the number of enterprises over time
(befare and after the era of structural adjustment policies) since the
mortality rates of informal sector enterprises at different points in time
are unknown. But since there are no clear a priori reasons to believe that
mortality rates differ markedly across either sectors or the rural/urban
divide, a table like this may allow us to make meaningful inferences with
respect to structural changes in informal sector activities. One question
which springs to mind, among others, is whether there was any inherent
bias in favour of trade among informal sector activities as a result of struc-
tural adjustment.

Collapsing the data table


To verify this hypothesis it is useful to collapse the data table to function
on this particular question. We can therefore introduce a simple trade/non-
trade dichotomous variable which will feature as dependent variable in
our subsequent analysis. Our main interest is to check whether the period
after structural adjustment policies were initiated brought about a shift in
favour of trading activities. Furthermore, our interest is to see whether
the pattem of change, if any, was similar in both urban and rural areas.
Table 9.3 summarises the data organised along three dichotomous cate-
gorical variables.

Logit modelling
How do logits come into our analysis? As we have seen, a logit is the
logarithm of the odds. In this example, we consider the odds in favour of
trading activities since the dichotomous 'trade/non-trade' variable is our
dependent variable. We therefore simplify Table 9.3 by introducing the
logarithms of the odds in favour of trade explicitly as the dependent vari-
able (Table 9.4).
How do we interpret this table? First start with the logits themselves.
For example, the logit in favour of trade for urban enterprises initiated
befare structural adjustment equals -0.0787. To interpret this value it is
useful to work your way back to the corresponding odds and proportion.
In general, if L is the logit, the corresponding odds, O, is obtained as
follows:
Logit transformation, modelling, regression 309
Table 9.3 Informal sector enterprises by trade/non trade, age and location
Trade Non-trade Total
Age Urban Rural Urban Rural Urban Rural
Less than 5 years 308,312 341,017 134,214 268,059 442,526 609,076
5 or more years 96,881 179,995 104,811 353,934 201,692 533,929
Total 405,193 521,012 239,025 621,993 644,218 1,043,005

Table 9.4 Logits in favour of trade by age and Iocation


Age Urban Rural Difference of logits
Less than 5 years 0.832 0.241 0.591
5 or more years -0.079 -0.676 0.598
Difference of logits 0.910 0.917 -0.007

o= eL (9.9)
which in this example yields:
o _ e--0.0787
";35,urban -
= 0.924
To obtain the corresponding proportion, p, we proceed as follows. Since

0=-p-
1-p
it follows that
o (9.10)
p = 1 +o

Substituting (9.9) into (9.10), we obtain:


eL
p = 1 +é
Finally, dividing both the numerator and the denominator by eL, we obtain
the following logistic function:
1
p = 1 + e-L (9.11)

which expresses a proportion as a non-linear function of the value of its


corresponding logit. In our example we obtain the following result:
P-,,s,urban = 0.48
310 Econometrics for developing countries
Exercise 9.2
Calculate the odds and proportions for the three remaining cases in
Table 9.4.

Our interest, however, is to analyse how odds differ or change in different


circumstances depicted by the explanatory variables in our model. Does
rural or urban location matter or not in terms of conditions favouring the
prevalence of trading enterprises? Did the change in policy regime favour
the development of trading activities within the informal sector? To answer
these types of questions explicitly we make use of odds ratios. In general,
the odds ratio is a ratio of the odds defined with respect to the depen-
dent variable (a dichotomous variable) at different values of the
explanatory variables (which, in this example, are also dichotomous in
nature ). For example, the odds ratio in favour of trade with respect to
urban or rural location of enterprises started befare structural adjustment
is given by:
- ourban,'35 - 96881/104811 -
[lurban/ruraf,'35 - O - 179995/353934 - 1.81
rural,~5

which reveals the much stronger prevalence of trading activities in urban


as against rural areas for enterprises which are five years old or more. An
odds ratio equal to 1 would indicate that the odds are equal in both
circumstances. Here this is not the case and, therefore, location clearly
matters.

Exercise 9.3
Calculate the odds ratios in favour of trade:
1 with respect to urban and rural location of enterprises initiated after
structural adjustment policies;
2 with respect to policy regime (before/after structural adjustment poli-
cies) for urban areas; and
3 with respect to policy regime for rural areas.

Now, the important thing to note is that simple differences of logits are
in fact logarithms of odds ratios. This result is very convenient since it
greatly enhances our ability to make sense of cross-tabulations of logits,
as is done, for example, in Table 9.4. To see this, we start with the odds
ratio defined as follows:

(9.12)

where 0 1 and 0 0 are two odds corresponding to different values of an


Logit transformation, modelling, regression 311
explanatory variable. Taking logarithms of both sides of equation (9.12)
yields:

In n = In ( ~J = In 0 1 - In 0 0

(9.13)
Conversely, if we computed both L 1 and L 0 , the odds ratio can be obtained
by calculating the anti-logarithm of their difference:
(9.14)
This simple result now explains why, in Table 9.4, we added a final row
and a final column of differences of the logits tabulated across dichoto-
mous explanatory variables. Take, far example, the difference 0.9104
obtained by subtracting the logit in favour of trade far urban enterprises
of five years old or more from the logit of similar enterprises less than
five years old. The odds ratio far urban areas in favour of trade as a result
of the shift in policy regime is then obtained as fallows:
í\
~L<S/;;,5,urban
-
-
eD.9104 -
-
2 •485

which indicates a strong swing in favour of trade in urban areas. In fact,


the column difference far rural areas yields a very similar result: an odds
ratio equal to 2.50. Consequently, the swing in favour of trade was very
similar in terms of changing odds in rural as well as urban areas.

Exercise 9.4
Using Table 9.4, show that the odds ratios between rural and urban loca-
tion in favour of trade were very similar far enterprises started befare
and after structural adjustment.

Interaction effect
But what about the computed value in the bottom-right comer of Table
9.4? This value is obtained by taking the difference of differences of logits.
Therefare, it involves differencing twice. Note, however, that it does not
matter whether we take the difference between row-differences or between
column-difference: both yield the same result. But how do we interpret
such difference of differences? The interpretation is quite straightfarward
notwithstanding the apparent complexity of differencing twice. The differ-
ence between two logits is the logarithm of an odds ratio; differencing
one more time yields the logarithm of the ratio of odds ratios. What this
measures is the interaction effect of both explanatory variables on the
dependent variable.
312 Econometrics for developing countries
In Table 9.4 we already noted that the column differences (respectively,
0.9104 and 0.9169) are very similar in magnitude. Consequently, the odds
ratio in favour of trade as a result of a change in policy regime is very
similar for both urban and rural areas. In other words, rural and urban
location does not interact with the change in policy regime (measured by
the age variable) as far as their impact on the prevalence of trading activ-
ities within the Tanzanian informal sector is concerned. The interaction
effect is measured by the logarithm of the ratio of the odds ratios. To
obtain the latter ratio, we only need to take the antilogarithm of the differ-
ence of differences between the logits in the table, as follows:

fl<S/;;oS µrban = e--D.0065 = 0.994


[2<5/;;oS ,rural

a result which confirms that the changing odds in favour of trade as a


result of the change in policy regime was pretty much alike in urban and
in rural settings.

Logit regression with a saturated model


The exercise we have just completed was an exercise in logit modelling.
It featured a qualitative dependent variable and two explanatory vari-
ables, all dichotomous in nature. The distinctive feature of this type of
modelling is that we seek to investigate how the probability of a partic-
ular characteristic being present (i.e. the enterprise is a trading enterprise)
varies with circumstances as depicted by the qualitative explanatory vari-
ables. To assess such changes in probabilities we use the logarithms of
odds ratios in favour of this particular characteristic. To show both the
similarities and differences with the regression models employed in earlier
chapters, it is instructive to work our exercise making explicit use of
dummy variables. Therefore, let T = l , if the enterprise is in trade, restau-
rant or hotel sector,= O, otherwise; U= l, if the enterprise is urban-based,
= O, otherwise; A = 1 , if the age of the enterprise is less than five years,
= O, otherwise. Table 9.4 can now be rewritten in a single equation which
seeks to explain the variation in the logits in favour of trade, LT, as follows:
LT = -0.68 + 0.60 u+ 0.92 A - 0.01 (A ·U) (9.15)
where A· U is the interaction dummy obtained by multiplying both
explanatory dummy variables. Equation (9.15) is in effect a logit regres-
sion in the context of a saturated model (meaning that the number of
estimated coefficients is exactly equal to the number of observations). In
fact, in this case, the interaction term can conveniently be dropped since,
as shown above, its coefficient indicates that the ratio of odd ratios is
clase to one. A more parsimonious model can thus be obtained by
regressing the logits in favour of trade against both dummy variables. But
Logit transformation, modelling, regression 313
this would lead us to logit regression, a topic we shall now discuss in
greater detail.

9.4 THE LINEAR PROBABILITY MODEL VERSUS LOGIT


REGRESSION
We saw in previous chapters that linear regression essentially implies
averaging Y for given values of the Xs. This makes good sense when the
dependent variable, Y, in the analysis is a measurement variable. But what
happens when the variable we are interested in is a qualitative dichoto-
mous variable which, as we have seen, can best be depicted by a dummy
variable which equals 1 when a particular characteristic is present and
zero otherwise? For any given observation, therefore, the characteristic
is either present (Y = 1) or absent (Y = O). In this case averaging <loes
not seem to make much sense. However, if you average a dummy variable
which takes the values one or zero across a sample of observations, the
result will give you the proportion of observations in the sample which
have the particular characteristic. In the context of random sampling this
proportion estimates the probability of encountering this characteristic in
the population at large. In other words, averaging in this context <loes not
tell us something about the average value a dummy variable assumes
(since it can only be either 1 or zero but nothing in between), but rather
it tells something about the probability that the dummy will equal l.
Consequently, when a qualitative dichotomous variable features as the
dependent variable in our analysis we seek to find those factors (explana-
tory variables) which determine the probability that the characteristic this
qualitative variable depicts is present in the population.

The linear probability model


Given that averaging a binary dummy variable, Y, gives us an estimator of
the probability that Y equals 1, it seems logical to extend the application
of linear regression to cases with a qualitative dependent variable. This is
indeed what is done in the context of the linear probability model. A sim-
ple example may help to illustrate this type of application. In Chapter 8,
we investigated the variation in income of Indian workers in relation to age,
sex and education. But what if our main interest is to see what determines
the prevalence of permanent jobs as against temporary or casual work?
In this case we end up with a qualitative dependent variable, F, which
equals 1 if a worker has a fixed job and zero otherwise. To keep the analy-
sis simple we shall limit the example to the bivariate case by investigating
the effect of age on the prevalence of permanent jobs. The simple regres-
sion of F; on A;, age, yields the following regression results:
F; = 0.058 + O.OllA;
(0.084) (0.0024) (9.16)
314 Econometrics far developing countries
R2 = 0.075, No. of observations = 261
Perhaps you find the R 2 disappointingly low. Keep in mind, however, that
we are dealing with a large sample of cross-section data. Moreover,
remember that our dependent variable is a dummy variable which can
only take on the values 1 or zero. Now, since the constant term is not
significantly different from zero, we can drop it from the equation to
obtain the following regression through the origin:
F; = 0.0127 A; R2 = 0.46
(0.0009) (9.17)
The slope coefficient did not change very much, but the R 2 is now much
higher. Keep in mind, however, that the R 2 in this regression is not compa-
rable to that in the regression with a constant term.

Exercise 9.5
Can you explain why the R 2 of the regression without constant term turns
out to be much higher than that of the regression with constant term,
although (in fact) the residual sums of squares of both regressions turned
out to be very close together? (Hint: Keep in mind that in a regression
through the origin the total sum of squares of the regression is not the
sum of squared deviations from the mean of the dependent variable (as
is the case when the regression features a constant term), but the sum of
squared values of the dependent variable, which in this case is a dummy
variable.)

The scatter plot with this regression line is shown in Figure 9.3. This plot
looks quite different from a scatter plot where both variables are measure-
ment variables, since all data points have a Y value 1 or O. The problem
with this plot, however, is that with a large sample it is quite possible that
different data points end up being superimposed on one another because
many of the workers concerned are likely to have the same age (measured
in years). To be able to see the distribution of the data points, therefore,
it is advisable to jitter the points a bit so that the thickness of the scatter
comes to the fore, as in Figure 9.3. (If your computer package will not
do this automatically then add a small random number to your dummy
variable for the purpose of drawing the graph.)
The regression line predicts the probabilities that the F;s equal 1 for a
given age. Hence, at age 15, our regression line predicts that the proba-
bility of encountering a worker with a permanent job is about 0.19, while
at age 55 the corresponding probability has risen to 0.70. This is why this
type of regression is called the linear probability model: it expresses the
probability that the dependent variable equals 1 for different values of the
explanatory variable(s). In this case, age clearly matters whether a worker
Logit transformation, modelling, regression 315

e:
Q)
e
C1l
E
(¡;
a...

15 20 25 30 35 40 45 50 55 60 65 70 75
Age
STaTa™

Figure 9.3 Scatter plot of F on A with regression line

is in permanent employment or not. The importance of the age variable


can also be shown by using the more familiar comparative box plots of
the age distribution for different values of the dependent variable.
At first sight, it appears that the linear probability model is quite appro-
priate to deal with a qualitative dependent variable. And, in fact, in this
example it does a good job. There are, however, two problems - one
minor, one major - with this model. The lesser of the two problems is
that the error terms in the model inevitably are heteroscedastics as shown
in Box 9.1. This problem can easily be remedied using a two-step proce-
dure involving weighted least squares (see Box 9.1). Using weighted
regression, our regression line through the origin now becomes:
S; = 0.013 A; R2 = 0.40
(0.001) (9.21)
which yields a similar result as in the case of ordinary least squares.
The second problem inherent in the linear probability model is far
more serious. By definition, a probability is bounded between O and l.
Consequently, the linear probability model only makes sense if the esti-
mated probabilities for the relevant range of X values stay within these
boundaries. But nothing guarantees that this will be the case. In our
example we were fortunate enough that the estimated probabilities for
316 Econometrics far developing countries

Box 9.1 Heteroscedasticity in the linear probability model


Consider the following general bivariate linear probability model:

Y¡ = ~1 + ~2 X¡ + E¡ (9.18)

where Y; is a dummy variable. Let P; be the probability that Y; = 1


and, hence, (1 - P;) is the probability that Y; = O since Y is a
binary variable. Consequently, E(Y;) = 0(1 - P;) + lP; = P; (the math-
ematical expectation of Y; equals the probability that Y; equals 1).
Hence:
E(Y; \ XJ = P; = ~ 1 + ~ 2 X; (9.19)
Now, since Y¡ can only assume the values O or 1, the resulting E; for
given X; can also only assume two outcomes:

Probability
o
1

Therefore, E¡ is also a binary variable with mean O and variance


P;(l - P;), the variance of a binomial variable with only one obser-
vation. Since P¡ depends on the value of X; it follows that the
different error terms corresponding to different values of X; will
have different variances. Consequently, the error terms will be
heteroscedastic.
The presence of heteroscedasticity leaves the estimator unbiased
but not efficient. It is preferable, therefore, to switch to weighted
least squares in arder to render the error term homoscedastic, as
shown in Chapter 7. The easiest way to do this is by using trans-
formed variables. This involves a two-step procedure, as suggested
by Goldberger (1964: 248-50):
Step 1 Estimate equation (9.18) with ordinary least squares and
compute the predicted Y; values to calculate the estimated
variances Y; (1 - Y;), the square root of which gives us our
weights, W¡.
Step 2 Estimate the regression:

(9.20)

with ordinary least squares without constant term.


Logit transformation, modelling, regression 317
70

60

50

40

30

20
Manual worker Operator
stata"'

Figure 9.4 Age distribution by worker category

given age within the age range from 10 to 70 years were all less or equal
to 1, or more than or equal to O. But this does not always happen, as can
be shown with a similar example taken from another part of the world
and in a different context.
This example concerns the selection of machine operators from a pool
of casual manual workers in Maputo harbour at the time of independence.
The problem emerged when, in Mozambique, many Portuguese settlers
who previously occupied most of the skilled jobs left the country in large
numbers after independence in 1975. The result was a grave shortage of
skilled labour in Maputo harbour, mainly of machine operators. A solu-
tion was sought by accelerated training and upgrading of workers selected
from a large pool of casual manual labourers. The age of the worker
appears to have been an important criterion for selecting trainee opera-
tors from among the casual workforce. To verify this hypothesis a stratified
random sample was taken of these two groups of workers: newly trained
operators and the remaining casual labour force.
As in the previous example, we have a dependent dummy variable, O,
which equals 1 if the worker is an operator and O otherwise (indicating
that the worker is a manual labourer). As befare, the explanatory vari-
able in the model is the age of the worker in years. The hypothesis we
seek to test is whether selection favoured younger workers. In fact, the
comparative box plot of the age of worker by type of work reveals that
318 Econometrics far developíng countries

-
~
Q)
c.
o

o ++
*

20 30 40 50 60 70
Age

Figure 9.5 Regression of O on age: linear probability model

+ ..,.,. +~+;ft;;*' ~+ ++ .... +# + +

o ++

20 30 40 50 60 70
Age

Figure 9.6 Linear probability model and logit regression: Maputo worker data
Logit transformation, modelling, regression 319
age matters, as shown in Figure 9.5. However, since age clearly is the
explanatory variable in this case we prefer a model which features the
type of worker as qualitative dependent variable. Estimating the linear
probability model with ordinary least squares yields the following results:
O; = 1.158 - 0.0204 A;
(0.187) (0.0045) (9.22)
R2 = 0.14; No. of observations = 129
a regression which confirms our hypothesis that the selection procedure
for upgrading workers favoured younger workers. But this regression is
in fact quite problematic, as can be seen from Figure 9.6 which depicts
the scatter plot with the corresponding regression line.
The problem is that our regression predicts negative probabilities from
the age of 57 years and above. A worker who was about 25 years old at
the time of independence had an estimated probability of 0.65 while a
worker of 56.5 years old would not be selected (zero probability). Above
the latter age, the probabilities predicted by this regression turn negative.
A further problem is that we can no longer apply the two-step procedure
of weighted regression to correct for heteroscedasticity either, since sorne
of our weights turn out to be negative.

Comparing Iogit regression with the linear probability model


How do we cope with this problem? One solution is to truncate the linear
probability model by ignoring the regression line whenever it cuts below
zero (or above one) and imputing a probability equal to zero (or one)
to the X ranges where such applies. Hence, in our example, we could
conclude that workers above the age of 56 had zero probability of being
selected as operators. This is a reasonable but not particularly elegant
solution to the problem. But what if we were to use logits instead of esti-
mated probabilities on the left-hand side of our model? Logits range from
minus to plus infinity and can always be reconverted into the corres-
ponding probabilities which will lie within the boundaries zero and one.
Logit regression helps us to do this: it features a logit as a dependent vari-
able which is assumed to vary linearly with the explanatory variable(s).
The mechanics of logit regression (which can be quite complex since it
involves non-linear estimation techniques) will be dealt with in the next
section. At this point we want to point out that if logits are linearly related
with the explanatory variable(s), it follows that the (estimated) probabil-
ities will be a non-linear function of the explanatory variable(s) since
equation (9.11) tells us that probabilities vary non-linearly with logits.
Figure 9.6. compares the regression line obtained with the linear prob-
ability model and the logit regression curve along with the scatter plot of
observations of the dummy dependent variable on the category of workers
320 Econometrics far developing countries

E
Q)
e
ro
E
(¡;
c...

10 20 30 40 50 60 70 80
Age

Figure 9. 7 Linear probability model and logit regression: ludian worker data

in Maputo harbour. The curve of estimated probabilities obtained through


logit regression avoids cutting below O and lends itself to easier interpre-
tation for this reason. The graph also shows that logit regression and the
linear probability model yield quite different estimates of the probabili-
ties concerned.
By contrast, Figure 9.7 gives us a similar comparison between logit
regression and the application of the linear probability model in the case
of our first example: the prevalence of permanent jobs as a function of
age for Indian workers. Here both models do equally well.

Exercise 9.6
Using the data set INDIA (Indian worker data) estímate the linear prob-
ability model between F and the age of worker, respectively, for all data
points, for male workers and for female workers. Construct the corres-
ponding scatter plots with regression line.

9.5 ESTIMATION AND HYPOTHESIS TESTING IN


LOGIT REGRESSION
In logit regression we are dealing with a qualitative dependent variable,
Y¡, which equals 1 when a particular characteristic is present in observation
Logit transformation, modelling, regression 321
i (i = 1 ... n ), or zero otherwise. Our interest is to predict the conditional
probability, P;, that Y; equals 1 for given values of the explanatory vari-
ables. We assume that the logit, L¡, is a linear function of the explanatory
variables:
(9.23)
and

(9.24)

which shows that the conditional probability is a non-linear function of


the explanatory variables. Our interest is to estimate the unknown coef-
ficients, 13; (i = 1 ... m), of the probability model (9.23). Once we obtain
such estimates we can calculate the predicted logits from equation (9.23)
and use equation (9.24) to obtain the predicted probabilities that Y; = 1
for given X; values.
At first sight it may appear that logit regression can be done easily with
ordinary least squares by regressing a transformed dependent variable,
the logit, on the explanatory variables. However, the problem is not as
easy as all that. Much depends on the nature of our data: more particu-
larly, on whether they are grouped or not. If our data are not grouped,
each observation involves an individual case for which Y; is either 1 or O.
The estimated probability for this case would thus be equal to 1 or O as
well. Consequently, the logit transformation will not work since the logits
corresponding to probabilities 1 and O are respectively plus and minus
infinity. It is not possible, therefore, to estimate equation (9.23) with
ordinary least squares since we cannot apply the appropriate logit trans-
formation to the individual cases where Y; is either 1 or O. With grouped
data, however, it is generally possible to apply the logit transformation
since each group will normally contain both cases with and without the
specific characteristic. This is the reason why we could apply the logit
transformation to our examples in sections 9.2 and 9.3.

Logit regression with individual case data


Consider first the case where each observation refers toan individual unit.
To estimate the logit model in this case requires more complex non-linear
estimation methods which are commonly based on maximum likelihood
rather than least squares.
Equations (9.23) and (9.24) show that P;, the conditional probability
that Y; = 1, is a non-linear function of the explanatory variables. Since
we are dealing with a binary dependent variable, P; is the probability that
Y; = 1 and, hence, (1 - P;) is the probability that Y; = O. To construct the
likelihood function, therefore, we note that the contribution of the ith
observation can be written as:
322 Econometrics far developing countries
pr (1 - p J(l-Y;)

and, in the case of random sampling where all observations are sampled
independently, the likelihood function will simply be the product of the
individual contributions, as fallows:

L = rrn
i=l
p.Y;
l
c1 - P.l
)1-Y; (9.25)

The principle of maximum likelihood entails that we choose those


values of the parameters of model (9.23) which maximise the likelihood
function. In practice, we maximise the logarithm of the likelihood func-
tion:
ln L = L[Y¡ ln P; + (1 - Y;) ln (1 - P;)] (9.26)
Substituting equation (9.23) into (9.26), and maximising equation (9.26)
with respect to the unknown parameters in the model will then yield a
system of non-linear normal equations (see, far example, Kmenta, 1986:
550-3) which can only be solved through iteration. Fortunately, nowadays
most statistical or econometric computer programs allow us to apply these
techniques routinely. What matters is to acquire the skill to interpret the
results of logit regressions in applied work.
To do this, take once more our earlier example on the prevalence of
permanent jobs among lndian urban workers. To broaden the scope of
our analysis we shall also include dummy variables on gender and on
educational attainment along with the age variable as possible determi-
nants of the probability that a worker has a permanent job. The results
we obtained are listed in Table 9.5.
The variables in Table 9.5 are defined as fallows:
F = qualitative dependent variable, = 1 if worker has permanent (fixed)
job, = O otherwise A = age of worker in years
DG = 1 far women, = O far men
D2 = 1 if educational attainment equals primary school, = O otherwise
D3 = 1 if educational attainment equals secondary school, = O otherwise

Table 9.5 Logit estimates


F Coefficient Standard error t-statistic
A 0.01 0.01 4.49
DG 0.14 0.35 0.41
D2 0.21 0.38 0.54
D3 1.12 0.35 3.17
D4 3.56 1.06 3.38
Intercept -2.61 0.49 -5.31
Log likelihood = -150.9
Number of observations = 261
2
X (s) = 52.91
Logit transformation, modelling, regression 323
D4 1 if educational attainment equals post-secondary schooling, = O
otherwise
To interpret the output of this estimation, let us start with the coeffi-
cients in the model. Obviously, as with linear regression, each coefficient
(apart from the constant term) depicts the change in the predicted logit
as a result of a unit change in the corresponding explanatory variable,
other things being equal.
The coefficient of the age variable, for example, tells us that the predicted
logit (i.e. log odds) in favoUT of fixed work changes by 0.0549 with every
increase in age by one year. However, most people find it hard to interpret
changes in logits. A more intuitive interpretation can be obtained by work-
ing with odds or odds ratios instead. In fact, the anti-logarithm of each slope
coefficient in a logit regression is an odds ratio at two different values, one
unit apart, of the corresponding explanatory variable. This interpretation
is very useful when dealing with dummy variables. In OUT example, work-
ers who completed secondary schooling as the highest level of educational
attainment face far better odds of having a permanent job than those with-
out schooling; the corresponding odds ratio equals:
n D3~Vno
-
schooling -
el.124
= 3.08

Exercise 9.7
Show that workers with post-secondary schooling are in an even more
advantaged position. To do this, compute the relevant odds ratio.

What is the odds ratio in favour of fixed work between workers with post-
secondary schooling and those with secondary schooling as the highest
educational attainment? This ratio can be computed as follows:
OD4=1
oD4=1/0 no schooling = 3.563
_e_ = l1 46
{}D4=l!D3=l = - Q
D3=1
oD3=1/0 no schooling e
1.124 .

If instead we are dealing with a measUTement variable such as OUT age


variable, for example, the antilogarithm of its slope coefficient gives us
the odds ratio corresponding to a unit change in the variable. In this case,
the odds ratio of a change in age by one year equals the antilogarithm of
0.05, or 1.06. Consequently, with each additional year of age, the odds
are again multiplied by this constant odds ratio. The odds ratio corres-
ponding to an age difference of ten years, therefore, will be equal to
(e0.0549)10, which yields 1.73.
Finally, we can always compute the predicted probabilities for given
values of the explanatory variables. We can do this using equation (9.11)
which tells us how to convert logits into the underlying probabilities. For
324 Econometrics for developing countries
example, a female worker, 25 years old, with only primary education
(D2 = 1) has a predicted probability to be in permanent employment
equal to:
L = -2.608 + 0.0549 (25)+ 0.143 (1)
+ 0.208 (1) + 1.124 (O) + 3.563 (O)
= --0.885

1
p = 1 + 1 / e--0.885 = 0.29

This type of calculation is quite cumbersome if done on a one-by-one


basis, but will prove very useful to design analytical graphs, called condi-
tional effect plots, to help us analyse logit regressions. This we do in the
next section.
So much for the interpretation of the coefficients in a logit regression.
But how do we handle hypothesis testing and assess the goodness of fit?
The t-statistics in logit regressions pose no special problems: they can be
dealt with much as you would do in linear regressions. In our example
we note that the age variable and the educational dummies corresponding
to secondary and post-secondary education along with the constant
term turn out to be statistically significant at the 5 per cent probability
level. The gender dummy and the dummy for completion of primary
schooling as highest level of educational attainment turn out to be not
statistically significant from O. Note that a zero coefficient in a logit
regression implies an odds ratio equal to e0 = 1, corresponding to a unit
change in the variable concerned. An odds ratio equal to 1 means that
the change in the variable concerned does not change the odds. As with
linear regression, however, two or more insignificant t-statistics in a logit
regression does not allow us to drop the corresponding variables jointly
from the regression. To do this we need a statistic akin to the F-test in
linear regression. In logit regression we use a chi-square statistic derived
from the likelihood ratio between restricted and unrestricted versions of
the model.
We illustrate how to do this with reference to Table 9.5, which listed
the results of our logit regression. This is the unrestricted version of our
model. The estimates for its coefficients were obtained by maximizing
the logarithm of the likelihood function (equation (9.26)). In this example
the estimated log likelihood equals -150.90. Any restrictions we choose
to impose on this model will obviously decrease the estimated log
likelihood. However, if these restrictions are reasonably valid in prac-
tice we would not expect the estimated log likelihood to decrease signif-
icantly. Consequently, the likelihood ratio of the restricted and
unrestricted models can serve as a useful indicator of the validity of our
restrictions.
Logit transformation, modelling, regression 325
Table 9.6 Logit estimates: restricted version of model
Coefficient Standard error t-statistic
A 0.05 0.01 4.48
D3 1.06 0.34 3.15
D4 3.54 1.05 3.37
Intercept -2.48 0.44 -5.59
Log likelihood = -151.10491
Number of observations = 261
2 = 52.50
X (3)

To test whether we can drop the dummy variables on gender and


primary education from the model, we formulate the restricted version of
the model subject to the joint hypothesis:
H0 : 13Dc = O and 13m = O
which yields the results listed in Table 9.6.
If we compare the restricted (R) and the unrestricted (U) versions of
the model we note that the coefficients of the retained variables did not
change much as a result of imposing the restrictions, while the estimated
likelihood decreased only slightly. The restricted model therefore appears
to be an acceptable, more parsimonious summary of the data. To test
more formally whether we can drop both dummy variables from the model
we make use of a chi-square statistic with two degrees of freedom (i.e.
the number of restrictions) based on the likelihood ratio, LR!Lu:
2 LR
X(2) = - 2 ln- = -2 (ln LR - lnLu)
Lu

= -2 (-151.10 + 150.900)
= 0.41 (9.29)

a value which leads us to accept the null hypothesis that the restricted
version of the model can be maintained.
As shown in Tables 9.5 and 9.6, the computer output of a logit regres-
sion will generally report a chi-square statistic which tests the hypothesis
whether all variables (apart from the constant term) can be dropped from
the regression. Consequently, the corresponding number of degrees of
freedom equals the number of variables included in the model (respec-
tively, five in the unrestricted version and three in the restricted variant
of the model). In both cases the null hypothesis that all variables can be
dropped is rejected by the data. Sorne computer programs may also report
a pseudo R 2 for logit regression. Following Aldrich and Nelson (1984,
quoted in Hamilton, 1992: 233) this pseudo R 2 is computed as:
326 Econometrics for developing countries

pseudo R 2 = · _X_ (9.30)


x+n
where the chi-square statistic tests the hypothesis whether all variables
but the intercept is zero and n is the sample size. In our example, the
pseudo R 2 is equal to 0.17 for the restricted model (and 0.17 for the un-
restricted version). Other versions of a pseudo R 2 are also in use but none
lends itself to the easy interpretation that can be given to the R 2 in linear
regression. For this reason little is gained by reporting a pseudo R 2 , but
it is useful to know that it is often reported by sorne statistical computer
software.

Logit regression with grouped data


So much for logit regression with 'individual case' data. But it also happens
frequently that we work with grouped or replicated data. For example,
most published survey data are grouped. Replicated data mean that there
are many observations on the dependent (dummy) variable for each
given combination of observations of the X variables. This normally
occurs when our X variables consist of a few dummy variables only. In
both cases (grouped or replicated data) it is possible to calculate observed
probabilities (proportions), odds and logits for the range of groupings
using equations (9.1), (9.3) and (9.5). Hence, for each group we calculate
the proportion of the number of successes (i.e. when the dummy depen-
dent variable equals 1) in the total number of cases and, subsequently,
calculate the implied odds and logits as observed in the groups. In case
a particular group contains only successes (or only failures), it is not
possible to obtain the calculated logit. In such cases, the group of obser-
vations is dropped from further analysis. It is important, however, to
reflect on whether such cases are purely accidental or could be expected
a priori.
Given that we can compute the observed logits across the range of
groupings (which have mixed successes and failures), it follows that we
can use ordinary least squares to estimate the logit regression:

LJ = 13 1 + 13 2 X¡ + ... + 13k Xk¡ + E¡ (j = 1 ... m) (9.31)


where, m = number of groupings; LJ = observed (calculated) logit for
group j.
For example, Table 9.4 gives us the observed logits in favour of the
trading sector for the Tanzanian informal sector in function of rural/urban
location and of the age of the enterprise (less than five years/more than
or equal to five years). Using dummy variables as defined in section 9.3,
the model in equation (9.15) can be estimated without an interaction effect
using ordinary least squares:
Logit transformation, modelling, regression 327
LJ = -0.675 + 0.594 U + 0.914 A
(-0.003) (0.003) (0.003) (9.32)
No. of observations = 4 R 2
= 1
a result which is virtually identical to the one obtained with the saturated
model in equation (9.15).
There is, however, a problem when using ordinary least squares to estí-
mate a logit regression with grouped data. The problem is that the error
term is heteroscedastic. This follows from the fact that, within each group
j, the number of successes follows a binomial distribution for which P¡
depends on the values of the explanatory variables. Furthermore, group
sizes will differ from group to group. It can then be shown (see, for
example, Kmenta, 1986: 551-2) that the error term in equation 9.31 has
a variance s;_¡ equal to

(9.33)

where n¡ is the number of observations in group j. If n¡ is sufficiently large,


the binomial distribution can be approximated by a normal distribution,
as follows:

(9.34)

To correct for heteroscedasticity, therefore, we compute the weights,


W¡,equal to the estimated group variances:
1
(9.35)

where P¡° is the observed (estimated) probability that the dependent


dummy variable equals 1 for group j, and re-estímate equation (9.31) using
weighted least squares

~W¡ L~ = /31 ~W¡ + /32 ~ X2¡ + ... + /3m ~W¡ Xmj (9.36)
In doing so, make sure not to introduce an explicit constant term when
carrying out regression (9.36).

9.6 GRAPHICS AND RESIDUAL ANALYSIS IN


LOGIT REGRESSION
This section is based on Hamilton's (1992: 217-47) excellent discussion of
diagnostic graphs in logit regression. As we have repeatedly argued in the
context of linear regression, analytical graphs help us to get more mileage
328 Econometrics far developing countries
out of our data. The same is true of logit regression. This section shows
you how to use two different types of graphs in logit regression: condi-
tional effect plots and poorness-of-fit plots. The former help us to
understand what logit modelling implies for predicted probabilities while
the latter is a diagnostic plot which enables us to spot which data points
fit the predicted pattern poorly.
In a bivariate logit regression with an explanatory measurement variable
it is always useful to produce a scatter plot of the binary dependent vari-
able against the explanatory variable along with the logistic curve of pre-
dicted probabilities. But how do we do this when we are dealing with a
multivariate logit regression? In this case, we are often interested in look-
ing at the curve of predicted probabilities in relation to one explanatory
measurement variable for given values of the other explanatory variables.
The resulting scatter plot and regression curve is a conditional effect plot.
Take, for example, the final logit regression arrived at in Table 9.6 after
weeding out irrelevant explanatory variables. We may be interested to
look at the relation between predicted probabilities of having a permanent
job and the age of the worker for given levels of educational attainment.
For example, setting D3 = O and D4 = O, we can calculate the predicted
probabilities of getting a permanent job for different ages in the case of
workers with educational attainment of primary-level schooling or less.
Plotting these predicted probabilities against the age variable gives us a

+ + ++-++..._. ~ ... ~+#'l"'-+:t:+ +tf' ++++ + *++

.9

.8

.7

E .6
(])
e
Cll
E .5
(i;
a.. .4

.3

.2

.1

10 20 30 40 50 60 70 80
Age
sra1a•M
Figure 9.8 Conditional effects plots
Logit transformation, modelling, regression 329
conditional effect plot, meaning that the plot is conditional upon both D3
and D4 being equal to O. Similarly, we could construct a conditional effect
plot for D3 = 1 and D4 = O and another one for D3 = O and D4 = l.
Figure 9.8 shows these three conditional effect plots all on one graph.
Each curve plots the conditional probabilities (given the level of educa-
tional attainment) against the age of the worker. Distances between the
curve for a given age show the importance of different levels of educa-
tional attainment. In this case, a higher level of educational attainment
increases the predicted probability of getting a permanent job. Notice, in
particular, that post-secondary education strongly boosts the probabilities
of getting permanent work.
Logit regressions, like linear regressions, can provide us with useful sum-
maries which give us insight into the data. But they can also mislead us
in much the same way as linear regressions do. Problems like multi-
collinearity, non-linearity, leverage and infiuence also arise in logit regres-
sion. In general, it is a good idea not to attempt logit regression when one
or more of the explanatory measurement variables is strongly asymmetric.
lt is best to transform this variable to avoid the problem of leverage and
infiuence. Furthermore, we should always supplement our logit regressions
with residual analysis, for which we may use Pearson residuals.
The Pearson residuals (Hamilton, 1992: 235-8) in logit regression are
not calculated with respect to each individual case, but with respect to
identical combinations of X values (or X patterns). If J denotes the
number of unique X patterns (J :s: n) and m¡ denotes the number of cases
with the jth X pattern, the Pearson residual is defined as:
{\

Y-mP.
(9.37)

{\

where P¡ is the predicted probability of cases with the jth X pattern and
Y¡ is the sum of observed Y values for cases with the jth X pattern.
At first sight, the Pearson residual may appear forbidding, but its
rationale is easy to grasp. In the numerator each group j (which may
include only one observation or more) refiects observations on Y, the
dependent dummy variable, for identical values on the explanatory
variables, X. Hence, M¡"P¡ = the predicted number of successes in this
group, and Y¡ = the actual (observed) number of successes, such that the
difference is the unexplained number of successes. The denominator is
just the standard deviation of the binomial distribution of the number of
successes.
The Pearson chi-square statistic (ibid.: 236) is the sum of squared
Pearson residuals:

(9.38)
330 Econometrics for developing countries
40
o

Q)

~ 30
O"
'i'
:.eu
en
"e:
o
~ 20
Q)
a.
.s:
Q)
Ol
e:
~ 10
ü
o

o
o 0.5
Predicted probability
s1a1a•M
Figure 9.9 Poorness of fit plot for logit regression

The closer Pearson's chi-square is to zero, the better the model fits the
data. In fact, in a saturated model Pearson's chi-square statistic will be
zero since all residuals equal O.
We can now obtain a measure of 'poorness-of-fit' of a particular X
pattern by calculating the change in Pearson's chi-square that results from
deleting all cases with this particular X pattern. Hence:
Ll X2p(j) = the change in Pearson's chi-square as a result
of dropping all cases with the jth X pattern (9.39)
A 'poorness-of-fit' plot can be obtained by plotting the change in Pearson's
chi-square as a result of deleting all cases with X pattern j against the
predicted probabilities, Pt Figure 9.9 provides such a plot for the logit
regression in Table 9.6.
When data points are situated near the horizontal axis in this type of plot
it means that the predicted probability does not differ much from the
observed probability and, hence, Pearson's chi-square will hardly change as
a result of deleting those cases which correspond to this particular X
pattern. A point which lies far above the horizontal axis fits the model
poorly. If it is situated on the right-hand side where the predicted proba-
bility equals one, it means that the actual data show that the particular char-
acteristic (Y = 1) is not (predominantly) present among cases with this
Logit transformation, modelling, regression 331
configuration of X values. An outlier on the left-hand side would indicate
that the data show that the characteristic is present while the model
predicts that it is not.
In fact, the outlier in Figure 9.9 corresponds to only one individual.
This was a man, 47 years old, with post-secondary education, who did not
have a permanent job. Our model predicts that the corresponding prob-
ability of being in a permanent job is very high, which explains the large
change in Pearson's chi-square provoked by deleting this observation from
the sample. Diagnostic plots such as that in Figure 9.8 allow us to spot
the unusual in a large data set. Once more, this illustrates the advantage
of using graphical tools in analysis.

Exercise 9.8
Use the data set INDIA (Indian worker data) to redo the estimation of
the logit model discussed above along with the conditional effect plots
and poorness-of-fit plot.

9.7 SUMMARY OF MAIN POINTS


1 When dealing with counts (frequencies) of a dichotomous qualitative
variable we can use proportions, odds and logits as convenient trans-
formations of the data. The logit transformation has the advantage
that it shows much better what happens in the tails of a distribution
(corresponding to very small or very large proportions ), while giving
a similar picture as proportions in the middle range.
2 Logit modelling with contingency data allows us to investigate the
relation between a dependent dichotomous variable and several
explanatory categorical variables. As such we separate the different
partial effects of each explanatory variable on the dependent variable
by translating the logits back into odds ratios. Similarly, interaction
effects can be obtained by twice (or more) differencing logits with
respect to two (or more) categorical variables.
3 The linear probability model seeks to model the probability that one
or another characteristic depicted by the dependent variable is present.
However, the model suffers from two defects: its residuals are
heteroscedastic and it may predict negative probabilities as well as
probabilities in excess of one.
4 Logit regression always predicts probabilities located between zero and
one. Logit regression can either be done with individual case data
(which requires an interactive ML estimation technique) or with
grouped data, using both dummy and measurement variables as regres-
sors. It extends the reach of regression techniques to deal with the
probability of sorne characteristic being present in the population as
depending upon a set of explanatory variables.
332 Econometrics far developing countries
5 The interpretation of the regression coefficient of a logit regression is
most conveniently done with reference to odds ratios. A zero coeffi-
cient implies an odds ratio equal to 1, which indicates that the odds
do not change as a result of a unit change in the explanatory variable.
6 Model choice through variable selection is possible in logit regression
by using the chi-square statistic in much the same way as we use the
F-statistic in linear regression.
7 It is advisable to use logit regression results jointly with graphical
displays such as the conditional effect plots and the poorness-of-fit plot
derived from Pearson's chi-square statistic based on Pearson's resid-
uals. The former plot renders interpretation of the results easier by
showing how different values of a particular explanatory variable affect
the probability of the characteristic being present along the range of
one of the measurement variables included in the model. The latter is
a diagnostic plot to spot outliers in logit regressions.
Part IV
Regression with time-series
data
This page intentionally left blank
10 Trends, spurious regressions and
transformations to stationarity

10.1 INTRODUCTION
In Part III we discussed problems specific to cross-section analysis. We
now turn our attention to time-series data. In such data the assumption
that the error terms from successive observations are uncorrelated is
frequently invalid; that is, investigation will find the residuals to be auto-
correlated. A likely cause of autocorrelation (which is itself discussed more
fully in Chapter 11) is that the series are non-stationary - a concept which
is defined in section 10.2.
Non-stationarity is a very serious matter: regression of one non-
stationary variable on another is very likely to yield impressive-seeming
regression results which are wholly spurious. Section 10.3 illustrates how
spurious regression may arise and section 10.4 presents formal tests for
stationarity. One possible means of avoiding spurious regression is to
transform the data so as to make them stationary - such transformations
are discussed in section 10.5. (The other solution to the problem is the
application of cointegration techniques which allow the estimation of non-
spurious regressions with non-stationary data; these techniques are
discussed in Chapter 12.) Section 10.6 summarises the main points from
the chapter.

10.2 STATIONARITY AND NON-STATIONARITY


A time series is a set of data connected in time with a definite ordering
given by the sequence in which the observations occurred. The time
ordering of the data matters a great deal since the moments of the distri-
bution of a time-series variable often change through time. That is, the
mean, variance, and so on of the underlying distribution from which an
observation is drawn are not constant, but depend on the point in time
at which the observation was made.
For example, in a growing economy the level of most macroeconomic
aggregates like production, consumption or the volume of trade will
increase over time. Figure 10.1 shows the consumer price index for Tanzania
336 Econometrics for developing countries
for the period 1966-92 and Figure 10.2 the rate of growth of real exports
for 1970-88 (the data for these figures are in the data file TANMON). Each
figure also shows a horizontal line drawn at the mean level for the period
as a whole. But, whilst to say that the average export growth rate over the
period as a whole was -5.0 per cent seems a useful summary of the data, it
means little to be told that the average CPI was 55.2 Why is this?
Recall that in Chapter 2 we modelled the mean of a variable as:
(10.1)
That is, the value of a variable is its mean value plus sorne random
error. Now, it <loes seem that the model described by equation (10.1) may
well fit the data for export growth - the time series <loes seem to fiuc-
tuate around the mean. But equation (10.1) is a wholly inappropriate
description of the CPI data - the variation of each observation from the
mean <loes not appear random at all. The lack of randomness of the error
term is - as is often the case - the result of the fact that equation (10.1)
is not the correct model. An alternative (but not the only alternative)
would be: 1
(10.2)
In equation (10.2) the expected value of Y 1 depends on t, in other words
the first moment is not time invariant. A variable, such as that given by
equation (10.1), whose moments are time invariant is said to be stationary.

320
300
280
260
o 240
o
íi
,.._ 220
a'.)

~ 200
x
Ql 180
"O
-~ 160
~
·g_ 140
Q; 120
E
::J
(/)
e: 100
o
() 80
60
40
20
o
1972 1984 1986 1988 1990 1992

Figure 10.1 Tanzanian consumer price index, 1966--92 (1987=100)


Trends, spurious regressions and stationarity 337
30

20

e:
Q)
10
u
(¡;
E,
.r:: o
~
e
C>
t -10
o
c.
X
Q)
(ij
Q)
a: -20

-30

-40 ~~-'---'----'-~-'---'----'-~-'---'-----'~-'----'---'~--'---'---''----'----'-~L....J

1970 1972 1974 1976 1978 1980 1982 1984 1986 1988

Figure 10.2 Tanzanian real export growth, 1970-88 (%)

If the moments are not time invariant then the variable is non-stationary.
The fact that many socioeconomic data are non-stationary has the very
serious consequence that regression results calculated with such series may
be spurious.

Invariance of series variance


We have just seen that the first moment of the CPI data systematically
increases across time. In addition, we often find that variation in the data
increases along with their level. Look again at Figure 10.1. Moving along
the scatter of points from left to right, the vertical distances between
successive points tend to widen. As the level goes up, so <loes the vari-
ance. Hence not only is the level of the CPI not stable over time, neither
is the variance. If we wish to transform the data so as to be stationary,
then both first and second moments must be rendered time-invariant. We
will see below how this may be done.
Here we have visually inspected for non-stationarity - a method which
is often sufficient to allow us to judge if a variable is stationary or not:
the Tanzanian CPI series is clearly non-stationary. In practice, formal tests
should be applied, and these are also discussed below. First, however, we
discuss why non-stationarity is such a problem.
338 Econometrics for developing countries
Exercise 10.1
Which of the following time series do you think would be non-stationary:
(a) nominal GDP; (b) real GDP; (c) growth of real GDP; (d) nominal
interest rates; (e) real interest rates; (f) external current account deficit
as a per cent of GDP; (g) life expectancy at birth; and (h) infant mortality
rate. Obtain data on each of these for the country of your choice (or use
the data in data file PERU) and plot them over time. (You will use these
data in several of the exercises in this chapter.) Do the graphs match your
expectations?

10.3 RANDOM WALKS AND SPURIOUS REGRESSION


How might we explain the rapid escalation in Tanzanian prices which
can be observed in Figure 10.1? One possible explanation is growth in
the money supply. As the best-known proponent of monetarism, Milton
Friedman, put it: 'inflation is always and everywhere a monetary phenom-
enon in the sense that it can be produced only be a more rapid increase
in the quantity of money than in output' (Friedman, 1970). Figure 10.3
shows the scatter plot of logged CPI against logged money supply (M2).
The figure also shows the fitted line from the regression, which is given
by the results (figures in parentheses are standard errors; DW is the
Durbin-Watson statistic to which we will refer below):
lnCP/1 = -5.69 + 0.91lnM2 1 R2 = 0.98
(0.25) (0.03) DW = 0.20 (10.3)
These results certainly seem to support the argument that monetary
growth has fuelled inflation. Yet it is quite possible that this apparent link
is wholly spurious. To see why this may be so we have to embark on a
digression concerning modelling trends in time series.

Modelling a random walk


The ordering of time series can matter since it contains the 'history' of
the variable. That is, the value of the variable in one period is an impor-
tant factor in determining the variable's value in the following period.
This facet of time series may be expressed in a number of time-series
models, the simplest of which is the autoregressive model with a single
lag, called AR(l). In this model the variable in time t is given by sorne
multiple of its value in period t - 1 plus an error term. Algebraically the
model is:
(10.4)
Spreadsheet simulations of equation (10.4) with different parameter
values are the best way to get a feel for the way an AR(l) model works.
First we set 13 1 = O and examine variations in 13 2 . If 132 = O then the series
Trends, spurious regressions and stationarity 339
6

*
5 *
*
*
'O
Q)
*
Ol 4 *
Ol
g
><
Q)
*
"O
3 *
.!: *
Q)
(.) *
"§_
•* *
(i; 2 *
E
::i
(/)
e:
o •*
ü
•* *

o
6 7 8 9 10 11 12 13 14
Money supply (logged)

Figure 10.3 Infiation and money supply: scatter plot of CPI against M2,
1966-92

1.0

0.5

-0.5

-1.0

-1.5 ~~~~~~~~~~~~~~~~~~~~~~~~~~~~

o 6 12 18 24 30 36 42 48 54 60

Figure 10.4 AR(l) process: 13 1 = O; 132 = O


340 Econometrics far developing countries
has no memory - the value of X in time t is given only by the error term
in that period and preceding values of X are of no relevance. This case
is shown in Figure 10.4. This graph therefore shows the 60 error terms
which will be used in the following simulations. (The graph contains
61 observations - .the first of these, X 0 , is given by an assumed initial
value, which is taken as 1 in this and subsequent simulations.) For future
reference, observe that only eight of the 31 observations from 30 to 60
are positive.
Figure 10.5 shows the simulation of the AR(l) model with 13 2 = 0.5, so
that the value of X is half of what it was in the preceding period plus a
random error. Compared with Figure 10.4, the series now appears to have
sorne 'memory' - that is, a tendency for the value in one period to be
closer to the value of the preceding period. Hence it can be seen that the
series takes a while to get away from its initial value of one. The succes-
sion of negative error terms in the second half of the series means that
X is likely to stay negative - there are only five positive observations from
periods 30-60.
The tendency for the series to stay where it was increases as the 13 2
coefficient increases towards unity. When 13 2 = 1 we have a special case
of the AR(l) process known as a random walk - so called because the
value of the variable in one period is equal to its value in the preceding
period plus a deviation by a random error. Figure 10.6 shows the simu-

o 5 10 15 20 25 30 35 40 45 50 55 60
Figure 10.5 AR(l) process: 13 1 = O; 13 2 = 0.5
Trends, spurious regressions and stationarity 341
lation of the random walk. Despite the fact that a random walk is gener-
ated by a succession of unrelated error terms the series displays apparent
trends. In this example the disproportionate number of negative errors
for observations 30-60 creates an apparent strong downward trend over
this period. Even a quite short run of error terms of the same sign will
send a random walk off on a seeming trend.
If the 13 2 coefficient exceeds unity then this 'trend' factor dominates the
series, since the variable in each period is its own past value times sorne
multiple. As may be seen from Figure 10.7, in which 13 2 = 1.1, the resulting
escalation in the X value soon drowns out the error term. The similarity
between this graph and that of the CPI for Tanzania (Figure 10.1) should
be readily apparent. It seems that the AR(l) model with a coefficient of
greater than unity would be a good representation of this data series.
A negative value of 13 2 results in oscillations, since the negative coeffi-
cient will reverse the sign of X in each successive period - though it will
sometimes happen that the error term will cancel out this reversal so that
two or more X values of the same sign may be observed. This cancelling
out effect is more likely if 13 2 is relatively small, resulting in a pattern of
oscillations which will appear more muted. In Figure 10.8 132 = -0.5 and
the 'jagged' nature of the time series is obvious.
What difference is made by non-zero values of 131 ? In the case when
132 = O then X will display random ftuctuations around a mean of 13 1. If

2.0

1.0

o.o
-1.0

-2.0

-3.0

-4.0

-5.0

--6.0

-7.0

-8.0
o 5 1o 15 20 25 30 35 40 45 50 55 60
Figure 10.6 AR(l) process: 13 1 = O; 13 2 = 1 - random walk
342 Econometrics for developing countries
300~~~~~~~~~~~~~~~~~~~~~~~~~~

250

200

150

100

50

o 5 10 15 20 25 30 35 40 45 50 55 60

Figure 10.7 AR(l) process: ~1 = O; ~2 = 1.1

o 5 10 15 20 25 30 35 40 45 50 55 60

Figure 10.8 AR(l) process: ~1 = O; ~2 = --0.5


Trends, spurious regressions and stationarity 343
132 has an absolute value of less than one then the series will fluctuate
around a mean of 13i/(l - 132), with apparent runs if 13 2 > O and in an
oscillatory manner if 132 < O. The former case is shown in Figure 10.9,
where the values of 13 1 (2) and 132 (0.3) give an expected value of X of
2.86. The graph of Tanzanian export growth (Figure 10.2) would seem
a candidate for being modelled as an AR(l) series with 13 1 < O and O <
13 2 < l. When 13 2 = 1 the series is known as a random walk with drift,
since the apparent trends may be reinforced by the addition of 13 1 each
period. This case is shown in Figure 10.10. The upward drift goes a long
way to cancelling out the strong downward trend that the random walk
showed in the later part of the series (Figure 10.6).

The algebra of stationarity


We see that the simple AR(l) model can generate a wide range of
different patterns of time series; these are summarised in Table 10.1. We
shall now demonstrate that a variable following an AR(l) process is
stationary if 11321 < 1, but not so otherwise. Repeated substitution into
equation (10.4) gives the following expression of X 1:
t-1 t-1
xt = 13~Xo + 131 I 13i + I 13i Et-i (10.5)
i=O i=O

o 5 10 15 20 25 30 35 40 45 50 55 60

Figure 10.9 AR(l) process: 131 = 2; 132 = 0.3


344 Econometrics for developing countries
?,-~~~~~~~~~~~~~~~~~~~~~~~~~

o 5 10 15 20 25 30 35 40 45 50 55 60

Figure 10.1 O AR(l) process: 13 1 = 0.2; 13 2 = 1 - random walk with drift

Interpretation of this expression is easiest by considering the three cases


of IB 2 1 < 1, (3 2 = 1 and JB 2 I > 1 separately.
First consider the case of a random walk:
t

B2 = 1 : xt = Xo + f31t + ¿ E¡
;~1
(10.6)

From this expression we can see that the value of X at time t depends on
three factors: (a) the initial value, X 0 ; (b) the amount of drift; and (c) the
sum of the current and all past error terms. Hence an error in one period
affects the value of X in all future periods: the series has a perfect memory.
This feature of a random walk may be contrasted with modelling X as
following a deterministic trend, that is:
(10.7)
Superficially equation (10.7) looks very like equation (10.6), but in fact
the structure of the error term is wholly different. In the trend, model X
follows the trend with random ftuctuations about that trend - but it always
returns to the trend as an error in one period is immediately 'forgotten'
in the subsequent period. By contrast, in a random walk the error is
embodied in all future values of X so that the series may stray a long way
from its underlying 'trend' value, and is not 'drawn back' to the trend as
Trends, spurious regressions and stationarity 345
Table 10.1 Series resulting from different values of 13 1 and 13 2 in AR{l) model
131 = o
13 2 = O X is just random error in X fluctuates in random manner
each period. No pattern around mean of 13 1
will be discernible
O < 132 < 1 X fluctuates around O with X fluctuates around mean of
'sorne memory' resulting 13 1/{l - 132) with sorne patterns
in short patterns
O > 13 2 > -1 X fluctuates around O in an X fluctuates around mean of
oscillatory manner 13 1/(1 - 132) in an oscillatory
manner
132 = 1 Random walk Random walk with drift
132 > 1 Explosive ( exponential) growth Explosive (exponen tia!) growth
132 < -1 Ever larger oscillations Ever larger oscillations

is a series generated by equation (10.7) - a random walk never forgets


an error. We saw an example of this in Figure 10.10 - despite the upward
drift the series declined in the second half of the period shown - and if
we were to run the model for another 1,000 periods we should have no
expectation that it would return to 'trend'.
The latter result runs contrary to the general beliefs of many econo-
mists, whose theories of long-run behaviour have been based on the notion
that variables such as real GDP follow a trend. Moreover the treatment
of the fluctuations around this trend (the business cycle) has been treated
as an independent area of enquiry. Evidence presented by Sims (1980)
that most majar macroeconomic aggregates for the USA could be
modelled as a random walk - so that the apparent trend and the cyclical
behaviour are the result of the same data generation process - struck
rather a blow to these traditional views. Subsequent work demonstrating
the bias and inconsistency of OLS using non-stationary series therefore
cast doubt on much published empirical work.
The expected value of a variable following a random walk is, however,
given only by its deterministic component. Taking expectations of equa-
tion (10.6) yields:

E(XJ = X0 + ~1t (10.8)


Recall that a non-stationary variable is defined as one whose moments
are time-dependent. From equation (10.8) it is clear that a random walk
with drift is non-stationary, since its first moment is a function of time.
However, the expected value of a random walk without drift (~ 1 = O) is
just the initial value, X 0 • This fact does not mean that such a variable is
stationary since its second moment is time dependent:
346 Econometrics for developing countries

E[X1- E(X1)]2 = E[~ EJ


=E[~~ E¡ E¡]= t a2 (10.9)

That is, the expected (squared) difference between X and its expected
value increases with time, which is to be expected as the 'history of errors'
will have a greater cumulative impact, allowing the series to wander away
from its underlying trend as time goes by.
Now consider the case of a variable with lr3 2I < l. In equation (10.5),
as t tends to infinity X can be written as:
/31 ~ .
l/321<1: xt ~ -_- + ~ /3 ~ €1-l as t ~ 00
1 /32 i=O (10.10)
The initial value is no longer relevant as the series forgets about events
too far in the past. Similarly the impact of a specific error term decreases
as it moves further into the past. The second component from equation
(10.5) is in this case equal to the expected value of ¡3 1/(1 - ¡3 2). The vari-
ance, as t ---+ > oo, is given by:
t-l t-l
E[Xt - E(X1)] 2 = ~ ~ /Íi/3~ E[ E¡_¡ E¡_¡]
i=O j=O

= a2(1 + fiz + IÍz + ...)


a2
(10.11)
=1-/iz
where we have used the fact that E( E;¡) = O for all i i= j and is equal
to cr2 for all i = j. A variable which may be described by an AR(l)
series with j[3 2 j < 1 is stationary - we have shown that its first and
second moments are independent of the time at which the observation
occurs.
Finally, equation (10.5) shows the explosive nature of a series in which
¡3 2 > l. Since ¡32 > 1 each of the three terms grows larger and larger as
t increases. Both first and second moments are time-dependent.

Significant regressions of independently generated random series


We have seen that when ¡3 1 = O and ¡3 2 = 1 then equation (10.4) describes
a random walk. As discussed earlier, a random walk process has 'a perfect
memory' so that the current value of X 1 is a function of all past disturb-
ances. Por this reason a random walk may display apparent trends, such
as can be seen in the X series shown in Figure 10.11 which we are to use
Trends, spurious regressions and stationarity 347

o 5 10 15 20 25 30 35 40 45 50 55 60

-- X --e- y

Figure 10.11 Two random walks

in the example below. In fact, a random walk is often called a stochastic


trend as distinct from a deterministic trend such as defined in equation
(10.7).
Figure 10.11 also shows Y 1 which was generated as a random walk:
Y1 = Y1_1 + E 11 1
E 1 - N(O, u 2) 't:lt (10.12)
Although the error term for the AR(l) process describing Y 1 has the same
distribution as the error term for the process for X 1 they are not the same
variable: the series of E will be different from the series E'.
It should be clear that the two series shown in Figure 10.11 have no
relationship with one another. Both are 'hypothetical' series based upan
independently generated series of random numbers. Yet if we regress Y 1
on X 1 we find a significant slope coefficient! The t-statistic is 3.64 (with
an R 2 = 0.18), compared to the critical value of 2.00 at the 5 per cent
level. OLS regression suggests that the independent random variable Y 1
is actually a function of X 1 - we know this is not true because we made
Y 1 ourselves, and it had nothing to do with X 1• This is surely a spurious
regression. Nor is this a freak result - repeating the experiment with a
different set of random numbers (and therefore different X 1s and Y1s) 120
times gave a significant relationship at the 10 per cent level between the
two variables in 96 (80 per cent) of cases - these results are given in the
frequency distribution shown in Table 10.2. In nearly two-thirds of cases
the relationship was significant at the 1 per cent level.
348 Econometrics for developing countries
Table 10.2 Frequency distribution of absolute t-statistic from 120 regressions of
unrelated random walks
Leve[(%) t-statistic Number of cases Per cent
t .;;:; 1.67 24 20.0
10 1.67 < t .;;:; 2.00 6 5.0
5 2.00 < t .;;:; 2.67 11 9.2
1 t > 2.67 79 65.8
120 100.0

The implications of this analysis are profound indeed. Regressing two


unrelated series on one another gives apparently significant results. Whilst
we have demonstrated the point by empirical example it has been shown
theoretically that application of OLS to non-stationary series yields biased
and inconsistent results.

lnflation and money supply revisited


So how do we know that the regression results reported in equation (10.3)
- showing a strong impact of money supply on the price level - are not
spurious? We don't. And they may well be. In fact the results we reported
give a strong indication that the regression probably is spurious. Spurious
regressions will often display bad residual autocorrelation. We shall deal
with this topic in the next chapter, when we shall meet the Durbin-Watson
(DW) statistic which provides a test for the presence of autocorrelation. For
the moment we may note a rule of thumb that if R 2 > DW there is a very
strong likelihood that the results are spurious. The regression of the CPI
on money supply yielded an R 2 of 0.96, but a DW of only 0.23. Hence we
should be very wary of accepting the results from equation (10.3) as proof
that infiation in Tanzania has been a largely monetary phenomenon.
Before interpreting time series regression results (ideally before con-
ducting any regressions) we must carry out two steps. First, we must check
whether the data are stationary or not. If they are not we must usually resort
to sorne technique other than OLS. Tests for stationarity are discussed in
section 10.4 and data transformations to stationarity in section 10.5.

Exercise 10.2
Use a spreadsheet to construct an AR(l) model and examine each of the
ten possible parameter combinations shown in Table 10.1.

Exercise 10.3
Using the graphs you drew for exercise 10.1, make an assessment of prob-
able parameter values for each series on the assumption that they may
be described by an AR(l) process.
Trends, spurious regressions and stationarity 349
Exercise 10.4
Use a spreadsheet to construct two random walks (generated by indepen-
dent error terms) and regress one on the other. Note down the resulting t-
statistic. Repeat a further 49 times. In what percentage of cases do you find
a significant result? What are the implications of your analysis for OLS
regressions using time-series data? (You may also do this exercise using an
econometric package with the facility to generate random numbers. If you
do use such a package you should also check your regression results to note
in which case R 2 > DW.)

Exercise 10.5
Can the problem of spurious regression arise m cross-section data?
Explain.

10.4 TESTING FOR STATIONARITY


A non-stationary series is one whose moments are not time-invariant.
Such a series can be expressed by the following general equation:
X( = 131 + l32t + 133Xt-1 + E¡ (10.13)
The nature of the time series described by equation (10.13) depends on
the parameter values - the most important distinction being between cases
which are a trend stationary process and those which are a difference
stationary process.

Trend stationary process versus difference stationary process


If 132 i= O and 113 31 < 1 then X follows a deterministic trend. The autore-
gressive component will mean that there may be short-run deviations, but
in the end the series will return to trend. 2 An example, with 13 1 = 10,
132 = 0.5 and 13 3 = 0.9, is shown in Figure 10.12. A series of this sort is
known as a trend stationary process (TSP), as the residuals from the
regression of X on a constant and a trend will be stationary. A series is
'detrended' by first regressing the variable on a time series and then using
the residuals from that regression in place of the variable.
On the other hand, if 13 1 i= O, 13 2 = O and 13 3 = 1 the series follows a
random walk with drift. The behaviour of such a series is very different
from that shown in Figure 10.12, since it displays no tendency to return to
trend. A random walk is known as a difference stationary process (DSP),
since the first differences will be stationary, where the first difference is
given by:
ax 1
= x 1
- x 1_1
= 131 + Er (10.14)
350 Econometrics far developing countries
?~~~~~~~~~~~~~~~~~~~~~~~~~~~

o 100 200 300 400 500 600 700 800 900 1000

Figure 10.12 Trend with stationary autoregressive component

The fact that many economic series appear to fallow a random walk
has two important implications. First, there do not exist long-run trend
values of these variables, from which they may depart but to which they
eventually return - a series of shocks can send the variable off on a wholly
different path far the rest of time. Second, as already mentioned, OLS
cannot be applied to non-stationary series.
The inapplicability of OLS is equally relevant to trend stationary
processes as to difference stationary processes. But if a variable were to
be a trend stationary process then OLS would be applicable to the
detrended series far that variable. But detrending a difference stationary
process, such as a random walk, <loes not make it stationary - so OLS is
still invalid. To apply regression techniques to a DSP, differencing must
be applied to make the series stationary. Differencing a TSP will also
result in a stationary series. That is, we can treat a TSP as if it were a
DSP and achieve valid results. But if we treat a DSP as a TSP we would
be led to the invalid application of OLS.
As mentioned, the first difference of a non-stationary variable may well
be stationary. Figure 10.13 shows the first differences far the lag of the
Tanzanian CPI and M2 (as we discuss on p. 360 these variables approxi-
mate the rate of infiation and monetary growth respectively). Whereas
Figure 10.1 showed the logged CPI to be clearly non-stationary the same
is not so obvious of its first differences.
Trends, spurious regressions and stationarity 351

0.4

0.3

0.2

0.1

o~~~~~~~~~~~~___.____.____.____.____.____.____.____.____.____.____.____.____.____.__,

1967 1970 1973 1976 1979 1982 1985 1988 1991

--- din (CPI) ---- din (M2)

Figure 10.13 First differences in logged Tanzanian CIP and M2, 1967-92

Order of integration (integrated variables)


If the first difference of a non-stationary variable is stationary, that vari-
able is said to be integrated of arder one, written as I(l). If second differ-
ences are required to achieve stationarity, then the variable is integrated of
arder two, and so on. Simple regressions should be carried out on variables
of the same arder of integration.
Saying that any time-series analysis should start by checking the arder
of integration of each variable is another way of repeating our general
exhortation to begin by looking at your data. You will find you can soon
make a reasonable guess of the arder of integration of a variable by
looking at its time path. What you are doing when you do this is to try
and match the picture you see against the different graphs generated by
the AR(l) process, shown in Figures 10.4-10.10 above.

Testing for non-stationarity (unit roots)


But it is also possible to test formally for stationarity by testing if the
variable may be described as a random walk, either with or without drift.
This procedure is also commonly called testing for a unit root as an I(l)
process has a unit root. If we find we cannot reject the hypothesis that
the variable is a unit root (e.g. follows a random walk) then we have
352 Econometrics far developing countries
found that the variable is non-stationary and must proceed accordingly.
It might seem that the test for stationarity is therefore a parameter test
on the AR(l) model given by equation (10.4). Specifically, we can test
the hypothesis that 13 1 = O and 132 = 1, in which case the series is a random
walk (i.e. non-stationary) if 13 2 = 1 and with drift if 13 1 i= O. Whilst this
approach seems intuitively appealing - and indeed is the basis of what we
do - it requires modification for the following three reasons:
1 Equation (10.4) is a more restricted version of the general model which
includes also a deterministic time trend (equation (10.13)), so we must
use the more general model as the basis for our initial testing.
2 It can be shown that if the series is non-stationary then the estimate
of 13 3 will be downward bias, so we are in danger of finding a series
to be stationary when in fact it is not. For this reason we cannot use
the standard t- or F-statistics when looking at critical values to conduct
significance tests. Alternative critica! values have been provided by
Dickey and Fuller (and another set provided by MacKinnon).
3 A modified version of equation (10.13) is used, called the augmented
Dickey-Fuller (ADF) test, which includes the lagged difference; the
inclusion of this term means that the critica! values are valid even if
there is residual autocorrelation.
The variety of appropriate test statistics are documented elsewhere
(e.g. Banerjee et al., 1993: ch. 4). Here we present a modified form of a
decision tree process suggested by Rolden and Perman (1994) to test
for stationarity. This process is rather more complex than the Dickey-
Fuller (or augmented Dickey-Fuller) test which the reader will commonly
come across. However, this test is made using a restricted version of the
general model (usually excluding both trend and drift) without having
tested those restrictions. If the restrictions are invalid then the critical
val ues are not applicable ( and if they are, then efficiency is lost if the
more general model is used). Hence, following the logic of Chapter 6,
the testing procedure should begin with the most general model; this is what
the decision tree does.
The tree structure, shown in Figure 10.14, allows us to move from a
more general to the more specific specifications of the equation, and indi-
cates the appropriate test-statistic at each stage. The critica! values are
given in the statistical tables on p. 480. Here we work through the deci-
sion tree, reporting results for the three Tanzanian series: logged CPI,
logged M2 and real export growth (XGROW). Note, however, that these
procedures are really only valid for large samples, so that their applica-
tion to the XGROW series is dubious.

Step 1 The first step is to estimate the most general model:


xr = 131 + l32t + 133X1-1 + 134U1-1 + e1,1 (10.15)
Estimate {lD.15) Accept y is a random walk without
drift
Yt=l1t+h+l1JYt-1+l,14AYt-1+Cr

and test restricted model

Ayi=l,146.Yt-l +e/

i.e. fio: ~1 =11 2=0 and ~3 =1


using calculated F against eritical
value <; 2

Reject Estimate Accept y is a random walk


with drift
Ayt=J11+l,14AYt-1+c¡"

and test against (10.15)


i.e. flo: ~=O and ~= 1
using calculated F agaínst cp 3

Reject Test fu=l from Accept y is random walk with


estimating (10.15) with deterministic trend (and
calculated t-statistic perhaps drift): an unlikely
against N(O, l) outcome for economic
series

Reject y has not got a unit root, so Accept y is stationary t<O:


usual t-statistics are valid to
test if 111 and l,12 are zero.
U se t-statistic to test ~=O
from equation (10.15)

Reject y is stationary t<O:y maybe


around a 1(2), repeat decisíon
deterministic tree using first
trend differences

Figure 10.14 Decision tree for testing stationarity


354 Econometrics far developing countries
Table 10.3 Application of decision tree for testing stationarity
ln(CPI) ln(M2) XGROW
(n = 25) (n = 25) (n = 17)
Unrestricted model RSSu 0.0692 0.1041 3,675
131 = 132 = o 133 = 1 RSSRI 0.1018 0.2056 8,426
F-stat 3.30 6.83 5.60
Accept Reject Reject
RSSR2 0.1399 8,420
F-stat 3.61 8.39
Accept Reject
133 = 1 Estima te -0.426
SE 0.357
t-stat -4.10
Reject
132 =o Estima te 0.225
SE 0.834
t-stat 0.270
Accept
Decision Random walk Random walk
without drift with drift Stationary

from which the RSS is the unrestricted sum of squares RSSu, shown for
each of the three series as the top line of Table 10.3. To examine if the
series is a random walk without drift we impose the restrictions ¡3 1 == ¡3 2
== O and ¡3 3 == 1, so that we estimate:

(10.16)
and use the resulting RSS (RSSRI in Table 10.3) to calculate the usual
F-test (see Chapter 6). The appropriate critica} value is given by the test-
statistic <1> 2.
In the case of ln(CPI) and ln(M2) the critica} value at the 5 per cent
level is 5.68, and is a bit larger than this for XGROW. (The 10 per cent
critica} value for a sample size of 25 is 4.67.) Hence we accept the null
hypothesis in the case of ln(CPI) and reject in the case of ln(M2). The
result for XGROW is clase to the critica} value. Given the weakness of
the test in smaller samples it is probably best to reject the null and proceed
to the next stage.
Accepting the null hypothesis indicates the variable to be a random
walk without drift. Not shown in the decision tree is the suggestion of
Holden and Pearson that we may now verify this result by imposing the
restriction that ¡3 2 == O in equation (10.15). The RSS from this equation is
then used to test the joint restriction ¡3 1 == O and ¡3 3 == 1 (for which the
appropriate RSS has already been derived, i.e. RSSRI) using the critica}
values given by <1> 1• In the case of ln(CPI), the RSS, once ¡3 2 is dropped
from equation (10.15), is 0.082, resulting in a calculated F-statistic of 1.77.
Trends, spurious regressions and stationarity 355
This value is below the critical value of 5.18 at the 5 per cent level, so
we accept the null hypothesis, thus confirming that ln(CPI) is a random
walk without drift.

Step 2 If the null from Step 1 is rejected we test instead whether the series
may be a random walk with drift. That is, 13 2 =O and 13 3 = 1, so we estimate:
fj.Xt = 131 + 134/j.Xt-l + E3,t (10.17)
The resulting RSS is shown as RSSR 2 in Table 10.3 (not for ln(CPI) which
has been established to be random walk without drift). Relaxing the
restriction that the intercept term is zero greatly reduces the RSS for the
ln(M2) series. Again the F-test is used, with the appropriate critical values
given by <1> 3 , which is 7.24 at the 5 per cent level for a sample size of 25.
Clearly we can accept the null for ln(M2), but have to reject for XGROW.
Logged money supply is non-stationary, that is, it follows a random walk
with drift.
If the null is accepted we may wish to verify the result by calculating
the t-test for the hypothesis that 13 3 = 1, using the critical value T3 •
Estimation of equation (10.15) for ln(M2) results in an estimate for 13 3 of
0.9088 with a standard error of 0.1074. The resulting t-statistic is -0.849,
compared to the critical value at 5 per cent of -0.80. This result casts
sorne doubt on the conclusion that ln(M2) is a random walk, but given
the marginal nature of the result ( the null hypothesis is easily acceptable
at the 10 per cent level) and the serious consequences of treating a random
walk as a non-stationary series, we had best proceed on the assumption
that ln(M2) is indeed a random walk with drift.

Step 3 If the null from Step 2 is rejected, the null hypothesis that
13 3 = 1 should now be tested with a t-statistic compared to a critical value
given by the standard normal tables. The estimated coefficient in the case
of XGROW is -0.4626, so the calculated t-statistic is -4.10. This value is,
in absolute terms, far greater than the N(0,1) critical value of 1.96 at the
5 per cent level, so we reject the null.

Step 4 If the null from Step 3 is rejected we know that the variable does
not have a unit root. We can test if it is stationary around a constant mean
or around a deterministic trend (i.e. is a trend stationary process, TSP).
Since there is no unit root, the usual t-tests are valid and we may use the
t-statistic, compared to the critical value from the usual t-tables to test the
null hypothesis that 132 = O. The estimated coefficient on the time trend
in equation (10.15) for the XGROW series is 0.225, so that the SE of
0.834 results in a calculated t-statistic of 0.27. This value is far below the
critical value of 3.16 at the 5 per cent level with 13 degrees of freedom,
so we accept the null, i.e. there is no deterministic trend.
356 Econometrics far developing countries
It therefore seems that XGROW is non-stationary (confirming our initial
visual impression), but it should be recalled that the application of these
techniques to a sample of less than 25 observations is not strictly valid.
The problem of small sample size is one that applied work, especially with
developing-country data, continually runs up against. Data quality is
poor, and where there are data, there are not many. Modern time-series
techniques require large samples, which we do not have, but traditional
techniques are simply invalid. There is no ready solution to this dilemma.
An important caveat to this procedure is that the variable may be 1(2)
(most likely with price series), in which cases the restricted model will
also be rejected at Steps 1 and 2. However, in this case the t-statistic in
Step 3 will be positive rather than negative (ie the estimated coefficient
is greater than one), so the variable is non-stationary. We may then check
the stationarity of the first differences.

Exercise 10.5
Replicate the results shown in Table 10.3.

Exercise 10.6
Test the stationarity of the data series you compiled for exercise 10.1.

Exercise 10.7
Use a spreadsheet to create two series, one following a trend with a
stochastic component (a TSP series) and the other a random walk with
drift (a DSP series). Alternatively you may use the data given in Appendix
10.1 for this exercise. Carry out an ADF test on: (a) each of the two series;
(b) the residuals from regressing each series on a constant and time;
and (c) the first differences of each series. Comment on your findings.

10.5 TRANSFORMATIONS TO STATIONARITY


To make a non-stationary series stationary it must be transformed in such
a way that its moments are no longer time-dependent. We have already
seen that a series which may be described by a random walk may be made
stationary by taking first differences; that is, it is a difference stationary
process. We pursue this point below, where we re-examine the link
between money and prices in Tanzania in the context of a difference
equation. But whilst differencing makes the mean (first moment) time
invariant, what of the variance? We first examine the problem of stabil-
ising the variance.
Trends, spurious regressions and stationarity 357
Stabilising variances
As just stated, to transform a time series to stationarity we need to stabilise
both its mean and its variance over time. This cannot be done with one
transformation only. In general, power transformations (using Tukey's
ladder of transformations: see section 3.5) are used to stabilise the
variance. The reason is that, most commonly, the variance of a time series
increases as its mean increases. The examples in Chapter 3 using cross-
section data showed that a power transformation can be very effective in
dealing with this pattern of covariance between level and spread. As we
shall see, the same is true with many time series.
As a rule, always start by stabilising the variance of time series befare
tackling the problem of its time-varying mean. How is this done in practice?
First of all, it is necessary to check whether or not a time series displays a
time-varying variance. Here is a simple method to do this. Slice the time
series up into roughly equal sections with between four and twelve observa-
tions in each section, depending on the total number of observations avail-
able. Take, for example, our time series of Tanzania's consumer price index
for the period 1966-92 (a total of 27 years ). We divide it up in to sub-samples:
1966-74; 1975-83; 1984-92 (hence, three slices ofnine observations). Figure
10.15 gives us a comparative display of the box plots for the three sub-
periods. This clearly shows that the spread increases as the level goes up; a
transformation is undoubtedly called for. But which transformation to use?

294.8

2.8
2 3
Sub-sample

Figure 10.15 Box plot for successive sub-samples of untransformed CPI data
358 Econometrics for developing countries
We could proceed by trial and error using the ladder of transforma-
tions, but section 7.4 above illustrated an easier way to determine which
transformation is likely to do the trick, which may be explained as follows.
Equation (7.4) was:
(10.18)
where ªx is the standard deviation of variable x, µx its mean and k a
constant indicating the data transformation required to stabilise the vari-
ance of the series. This equation suggests a practical method to check
which transformation is likely to be best. First, plot the logarithm of a
measure for spread against the logarithm of a measure for level (average)
for successive slices of a time series. Usually we use the interquartile range
and the median instead of the standard deviation and the mean because
the former are more robust. If the successive points (plotted accordingly
to this method) roughly approximate a straight line, the pattern of covari-
ance behaviour level and spread is stable through time: the lag-linear
model given by equation (10.18) is the appropriate one. Having plotted
this line, computing its slope indicates which transformation is likely to
do the trick, since this slope corresponds to the variable k.
Figure 10.16 plots the logarithms of the interquartile range for each
slice against the logarithm of its corresponding median (see Table 10.4).
The scatter of the three points shows a steady increase which can be

4.0

3.5

3.0

2.5
((

º 2.0

1.5

1.0
1.5 2.0 2.5 3.0 3.5 4.0 4.5
Median

Figure 10.16 Logged IQR against logged median for Tanzanian CPI
Trends, spurious regressions and stationarity 359
Table 10.4 Medians and interquartile ranges
Median Interquartile range
Value Lag Value Lag
1966--74 4.40 1.48 1.85 0.62
1975-83 11.90 2.48 13.00 2.56
1984-92 131.10 4.88 152.00 5.02

QL---1.-1..-1..-1..-1..-1..-1..-1..-1..-1..-1..-1..-1..-1..-1..--1.--1.--1.--1.--1.--1.--1.--1.--1.--1.--1.--1.-'

1966 1969 1972 1975 1978 1981 1984 1987 1990

Figure 10.17 Logged Tanzanian CPI, 1970-88

reasonably approximated by a straight line. To calculate its slope we can


use the upper and lower points:
k = 5.02 - 0.61 1.29 (10.19)
4.88 - 1.43
so that k = 1, suggesting that the logarithmic transformation is the right
one to stabilise the variance.
Figure 10.17 plots the logarithm of Tanzania's consumer price index
against time. As we can see from this graph, the vertical distances between
successive points no longer show the earlier tendency to increase over
time (although there is sorne variation, showing that the transformation
has not completely stabilised the variance ). In general, in economic
analysis it is the case that logarithms stabilise the variance, although, some-
times a series may be encountered far which the square root transfor-
mation is appropriate.
360 Econometrics far developing countries
Exercise 10.8
U se the procedure outlined above to determine the appropriate trans-
formation to stabilise the variance for each of the series you compiled for
exercise 10.1.

Stabilising the mean


Once we succeed in stabilising the variance through time, we can then
tackle the problem of a time-varying mean. A power transformation used
to stabilise the variance cannot <leal with stabilising the mean as well. The
reason is that power transformations alter the shape of the distribution
of the data but, at the same time, preserve their ordering. To stabilise the
mean we need a different type of transformation altogether.
Historically, the detrending of a time series was the common method
to stabilise its mean. A trend line (usually, a polynomial) was fitted to the
data, leaving residuals which, consequently, no longer displayed a time-
varying mean. The practice of detrending implicitly assumes that the series
is a TSP, i.e. follows a deterministic trend to which it will always revert. But
we have found that many series may be described as following a random
walk, and detrending such a series <loes not remove the problem of non-
stationarity (see exercise 10.7). To do this, differences must be taken.
We saw that both the CPI and M2 are non-stationary, so that regres-
sion of one on the other is not generally valid. (As discussed in Chapter
12, it will only be valid if the two variables are cointegrated.) The variances
of each variable can be stabilised by using logs. To stabilise the mean
we must take first differences. The first difference of the natural log of a
variable has the additional advantage of being approximately equal to that
variable's rate of growth, that is, the rate of inflation and monetary growth,
respectively (see Box 10.1).
Suppose a true relationship is given by:
(10.20)
then it follows that:
(10.21)
where w = Ei - Ei-J· That is, the same slope coefficient should be revealed
in both the levels and difference regressions (and the intercept in the
difference equation should be zero ).
Recall that equation (10.3), in which the logged CPI was regressed on
logged M2, showed an apparently strong relationship between the two
variables. Estimation of equation (10.21) with these data gives (standard
error in parentheses ):
iiln(CPI) = 0.79iiln(M2) R2 = -0.26
(0.08) DW = 1.16 (10.22)
Trends, spurious regressions and stationarity 361

Box 10.1 Logged differences and rates of growth


For small values of x the following approximation is true:
ln(l + x) = x
Therefore:
1
ln (X1) - ln (X1_ 1) = ln ( : ) = ln (1 + gx) = gx
t-1

where

That is, the first difference of the log of x is approximately equal to


the rate of growth of X.
As an exercise you could use a spreadsheet to examine the range
of values of gx over which this approximation is valid.

Although the t-statistic for the slope coefficient is high, the R 2 of -0.26
clearly indicates that we have invalidly omitted the intercept, which will
bias the estímate of the slope coefficient. This fact alone should cast doubt
on whether the regression of ln(CPI) on ln(M2) is the true model.
Including an intercept gives (standard error in parentheses):
dln(CPI) = 0.13 + 0.25dln(M2) R 2 = 0.06
(0.05) (0.21) DW = 0.74 (10.23)
The R 2 remains pitifully low. More importantly, the slope coefficient is now
only 0.25, which is far less than the 0.91 obtained in equation (10.3). The
growth of the money supply now appears to have no significant effect on
the rate of inflation. What happened? The regression between the levels of
both variables looks good, but as soon as we switch to regressing changes
in these levels on one another the results are disappointing. It appears,
therefore, that first differencing took the wind out of our sails. Why?
The problem líes in the regression between the levels of both time
series. In regression analysis, the total variation of a dependent variable,
Y, around its mean is in part explained by the variation in the indepen-
dent variable, X, around its mean, leaving the residual variation as the
remainder. If, as is the case in our example, both Y and X are stochastic
variables, we assume that they are jointly sampled from a bivariate normal
distribution. This implies that we assume that each variable (as well as
the error term) is distributed normally with a constant mean and a
constant variance. The regression line then explains the variation in Y
362 Econometrics far developing countries
around its constant mean in terms of the variation in X around its constant
mean. Hence, each Y1 ( or X 1) is assumed to be sampled from the same
distribution with a constant mean and variance. However, in our example
this clearly is not the case. The problem is that neither the dependent nor
the independent variable has a constant mean through time. Successive
observations of either of these variables appear to be drawn from different
distributions with progressively higher means.
Take another look at Figure 10.1. It seems far-fetched to assume that
this is a sample of observations drawn from a distribution with a constant
mean or variance. The log transformation we subsequently applied to this
variable in regression allows us to stabilise the variance, but obviously not
the level. The reason is that logarithms preserve order although they alter
the shape of the data.
Hence, the results we obtained by regressing the levels of both log trans-
formed variables may be as much due to the fact that each variable is
constantly on the move in terms of its level as to any real relationship
between them. This is what make regressions with time series so prone
to the danger of spurious correlations. In Chapter 12, on cointegration,
this issue will be taken up again. There we shall show that in sorne
instances it is possible to derive meaningful results by regressing time
series which are not stationary over time.
Why did differencing alter the picture so dramatically? The reason is
that taking first differences often yields a new time series which no longer
manifests a changing mean through time. Consequently, the assumptions
of ordinary least squares are more likely to be valid in practice. We can
now relate the variation in Y around its constant mean to the variation
in X around its constant mean. The regression, therefore, is much less
likely to produce spurious results. It follows that with time series it is
often useful to transform the original variables in a manner which stabilises
their means and variances through time. Regressions with such trans-
formed variables are less likely to fall prey to spurious correlations.
Indeed, we can use the fact that the regression in differences yields
different results to levels regression as a test of whether or not a regres-
sion is spurious.

An omitted variable version of the Plosser-Schwert-White


differencing test
Suppose our model is:
(10.24)
We are interested to test if the regression in differences yields the same
parameter estimates of the slope coefficients as the regression in levels.
It can be shown that this test is equivalent to testing the joint restriction
that -y 1 = -y 2 = O in the model:
Trends, spurious regressions and stationarity 363
(10.25)
where
zi,t = X1,1-1 + xi,1+1 (10.26)
The null hypothesis is that differencing gives insignificantly different
results, so the levels regression is not spurious.
The test can readily be generalised to the case of more right-hand side
variables. If one of these variables is the lagged dependent variable then
the corresponding Z variable is:
Zy,t = Yi,1-2 + Y;,1+1 (10.27)
with the other Z variables still defined as given by equation (10.26).
As an illustration we apply the test to the regression of logged Tanzanian
CPI on the logged money supply. Recall that differenced equation had a
negative R 2 , suggesting that we should use an intercept - which would
not be the case if the parameters of the differenced model corresponded
to those from the model in levels. Thus we suspect that the results may
well be spurious. We define Zas the sum of lagged and one-period forward
logged money supply and regress ln(CPI) on ln(M2) and Z. The results
are (t-statistics in parentheses):
ln(CPI) = -5.79 - 0.28ln(M2) + 0.60 Z R2 = 0.98
(-19.89) (-0.26) (1.22) (10.28)
Since we are testing the restriction on only one Z variable a t-test may
be applied (as it is equivalent to the F-test - see Chapter 6). The t-statistic
is 1.22, so we accept the null hypothesis that Z can be dropped from the
regression. That is, we find the results are not spurious. This finding does
not match with our expectations, and we should perhaps seek confirma-
tion from another test; this will be done in Chapter 12.

Exercise 10.9
Calculate the series of first differences for each of your series from exer-
cise 10.1, having first applied the appropriate power transformation as
determined in exercise 10.8. Conduct the ADF on each of the resulting
series. How do your results compare with those obtained in exercise 10.6?

10.6 SUMMARY OF MAIN POINTS


1 Time-series data have special properties deriving from their unique
ordering.
2 A non-stationary time series is one whose moments are not time-
invariant.
3 A convenient class of models for modelling time series are given by
the AR(l) process. With an autoregressive coefficient of one, the
364 Econometrics for developing countries
AR(l) model is a random walk (or random walk with drift if there is
an intercept term); such series are described as having a unit root.
A random walk is a non-stationary process.
4 Regression of one random walk on another is likely to yield a signif-
icant result even if the two series are totally unrelated to one another.
More generally, application of OLS to non-stationary series yields
biased and inconsistent results (except in the special case of the vari-
ables being cointegrated - see Chapter 12).
5 If a variable follows a trend stationary process it can be made
stationary by detrending. But most economic variables seem better
described by a difference stationary process - in which case differences
must be taken to stabilise the mean. A power transformation must
also first be applied in most cases to stabilise the variance.
6 Testing for a unit root is carried out by estimating an autoregressive
model, allowing for the possibility of an intercept and deterministic
trend. The procedure tests down to find the appropriate specification
and corresponding value of the coefficient on the lagged dependent
variable.

ADDITIONAL EXERCISE

Exercise 10.10
Generate 61 observations for a variable following a random walk without
drift. Regress the variable on its own lag 200 times, each time noting the
value of the estimated slope coefficient and the value of the t-statistic for
the test of the null that the slope is unity. Plot a histogram of the slope
coefficients. How often is the null rejected at the 5 and 10 per cent levels?
Why are these results a surprise? What do they tell you about testing for
stationarity?
Trends, spurious regressions and stationarity 365
APPENDIX 10.1:
GENERATED DSP AND TSP SERIES FOR EXERCISES

Time TSP DSP


o 0.00 0.00
1 5.93 15.61
2 15.58 19.48
3 36.29 31.51
4 41.11 38.13
5 39.00 70.76
6 43.87 100.26
7 52.82 131.19
8 60.31 99.90
9 59.55 96.26
10 73.03 116.62
11 88.84 137.85
12 93.81 144.28
13 106.88 128.05
14 102.19 150.66
15 102.96 119.05
16 111.96 122.41
17 117.91 109.55
18 126.25 121.91
19 131.59 139.14
20 138.90 137.43
21 144.81 118.56
22 157.17 119.76
23 168.70 135.14
24 180.55 137.35
25 176.80 158.92
26 179.94 187.65
27 189.05 170.25
28 198.35 187.97
29 203.92 199.00

NOTES
1 This specification is not appropriate, if only because the relationship between
the CPI and time is clearly nota linear one. More importantly, the series may
be a difference stationary process rather than a trend stationary process - this
distinction is discussed on p. 344--5 and 349-50.
2 The trend is given by the expected value, which is:

E(X) ~ f3i + t + !33 as t ~ oo


1 1 - {33 (1 - {33)2
11 Misspecification and
autocorrelation

11.1 INTRODUCTION
Autocorrelation (also called serial correlation) is the violation of the
assumption that E( E;E¡) = O. When the error in one period is related to
the error in another period then OLS is no longer BLUE. Moreover, the
R 2 may be overestimated, standard errors underestimated and t-statistics
overestimated. If the regressors include a lagged dependent variable then
OLS estimates are biased.
The presence of autocorrelation in the residuals of the estimated model
is, however, often a result of model misspecification, rather than 'genuine'
autocorrelation of the model error term. Recall that formal tests of the
property of the error term are carried out on the residuals. But, whilst
the error is a part of the data generation process, the residuals are a
product of our model specification. Hence testing for autocorrelation
should in the first instance be interpreted as a test for misspecification. A
range of techniques are available to detect autocorrelation. Here we will
present graphical methods, the runs test and the Durbin-Watson statistic,
we also define Durbin's h which should be used in the presence of a lagged
dependent variable.
This chapter is organised as follows. section 11.2 explains in more detail
what autocorrelation is and why it is a problem and section 11.3 considers
the various reasons why autocorrelation may be present. Formal tests for
autocorrelation are presented in section 11.4, and section 11.5 discusses
how to <leal with autocorrelation. Section 11.6 summarises the chapter.

11.2 WHAT IS AUTOCORRELATION ANO WHY IS IT A


PROBLEM?
The classical linear regression model assumes there is no serial correla-
tion - that is, a zero covariance between the error terms of different
observations. Hence, in the model:

(11.1)
Misspecification and autocorrelation 367
we are concerned here with the assumption that E( E1E1_,) = O. (We are
now using a t subscript as autocorrelation is a time series problem.) Put
more simply, because the error for one observation is large this <loes
not mean that the next one will be. Indeed, the fact that an error term
is positive should have no implications for whether the next term is
positive or negative.
To understand the implications of autocorrelation for residual plots and
the basis for tests for autocorrelation it is useful to spend sorne time
looking at the autoregressive (AR) model, introduced in Chapter 10. The
AR model is serially correlated by construction, so it is a good device for
seeing what such an autocorrelated error will look like in practice. But
you are not required to construct these AR models as a part of testing
for autocorrelation in the normal course of events.
Suppose that the error term, E, in the model of equation (11.1) is gener-
ated by an AR(l) process with a white noise error (v):
v - N(O, u~) Vt (11.2)
Figure 11.1 shows the case in which p = O, so that E is just equal to that
period's error term, v. 1 We know here, because we have generated the
data, that there is no serial correlation - each error term is indeed inde-
pendent of the others. How can this fact be seen in the residual plot? If
the different terms are independent then we should not see any patterns
in the data - for example, there should not be long runs of positives

1.0

0.5
e:
.Q
·¡¡:;
a.
w

-0.5

-1.0 ~~~~~~~~~~~~~~~~~~~~~~~~~

o 10 20 30 40 50
Period

Figure 11.1 No autocorrelation (p =O)


368 Econometrics far developing countries
followed by long runs of negatives. (A run is defined as successive values
of the same sign.) Equally, a positive value in one period should not imply
a negative value in the following period, so the plot should not be excep-
tionally jagged (i.e. have many short runs of positive and negative values).
Figure 11.1 seems to satisfy these conditions. But now look at Figures
11.2, generated using the same set of errors but with p = 0.7 and
Figure 11.3, again generated with the same errors but with p = -0.7. In
both of these cases a pattern may be observed. When p is positive then
the value of E is equal to sorne fraction of its previous value plus the error
term - so that a positive E in one period is likely to imply a positive one
in the following period (the next period's E will only be negative if there
is a quite large negative error). As can be seen in Figure 11.2, this model
produces a number of quite long runs of positive and negative values:
there are eight runs in all, four positive and four negative (a single posi-
tive or negative observation counts as a run). In Figure 11.3 by contrast
there are very many short runs: 32 in all. The deterministic part of the
AR(l) model with a negative value of p reverses the sign each period -
this reversa! may sometimes be cancelled out by the error term but is not
generally so.
What we see here is that when E is generated by an AR(l) model with
a positive coefficient, the time plot of the variable displays apparently
long runs of positive and negatives - this is called positive autocorrela-
tion (because it comes from a data generation process with a positive

2.0

1.5

1.0

0.5
e
.Q
·¡¡;
c.
o.o
w
-0.5

-1.0

-1.5

-2.0
o 10 20 30 40 50
Period

Figure 11.2 Positive autocorrelation (p = 0.7)


Misspecification and autocorrelation 369
1.5

1.0

0.5

e:
..Q
·¡¡;
a.
o.o
UJ

-0.5

-1.0

-1.5
o 10 20 30 40 50
Period

Figure 11.3 Negative autocorrelation (p = -0.7)

coefficient). With a negative coefficient there is negative autocorrelation,


which is shown by very short runs in the data (i.e. a jagged plot). Hence
a first test for autocorrelation is to look at the residual plot from the
regression to see if there appear to be too few (or too many) runs. A
more formal test for autocorrelation is the runs tests, which tests if there
are indeed too many or too few runs to be consistent with the null hypo-
thesis that the data are generated in a serially uncorrelated manner. The
other test we consider below, the well known Durbin-Watson test, is a
test, to see if the residuals are generated from an AR(l) model.
We said above that data generated by the AR model are serially cor-
related. More formally we may show (see Appendix 11.1) that:
2
E( e,z) = ____<!_,,___
1 -p2 (11.3)

(11.4)

Equation (11.4) shows that the ES generated by the AR(l) process are
indeed autocorrelated, since the covariance of any pair of ES is not zero
(though it tends to zero as the ES get further apart). However, equation
(11.3) shows that the variable is homoscedastic. 2
370 Econometrics far developing countries
Why is autocorrelation a problem?
The proof that LS estimators are BLUE uses the assumption that the error
term in the model is not autocorrelated. Violation of this assumption means
that the proof is no longer valid, and it is indeed the case that OLS is
no longer BLUE. 3 We do not, however, need this assumption of no auto-
correlation to prove unbiasedness, so the estimates remain unbiased
(except when the model contains a lagged dependent variable, as discussed
below).
In addition to the loss of efficiency, the residual variance no longer
provides an unbiased estímate of the error variance. Hence the attempt
to construct confidence intervals or to test hypotheses about the coeffi-
cients is made invalid, since the standard errors are no longer applicable.
When there is positive autocorrelation - which is the more common sort
in practice - then the estimates of the error variance have a downward
bias. Hence our confidence intervals are narrower than they should be
and the calculated t-statistics inflated, so that there is a danger that we
shall incorrectly reject the null that the variable has no significant impact.
Likewise the R 2 and related F-statistic are likely to be over-estimated.

Autocorrelation in the presence of a lagged dependent variable:


particularly bad news
If the model contains a lagged dependent variable then the OLS estima-
tors are not only not the most efficient, but they are also biased. Consider
the model:
(11.5)
where
(11.6)
Now, the proof of unbiasedness requires that there be no relationship
between the model regressors and the error term. In this case we require
that E(Y1_1e1) = O. But it is easy to show that this requirement is not met.
Lagging equation (11.5) by one period we get:
yt-1 = 131 + l32Xt-l + 133Yt-2 + Et-1 (11.7)
From which it follows that:

(11.8)

11.3 WHY DO WE GET AUTOCORRELATION?


It is very important to distinguish between genuine autocorrelation and
that which is a product of model specification. The traditional approach
Misspecification and autocorrelation 371
emphasises the former (sometimes exclusively so), whereas there are
very many cases in which the autocorrelation is in fact a symptom of
model misspecification. In this section we discuss these different pers-
pectives.

The traditional view


A common example of autocorrelation is an agricultura! production func-
tion. Suppose we regress crop output (Q) on the main determinants of
output - in this case a relative price variable (P) and fertilizer input (F).
The error term captures the random effects excluded from our model.
But a negative shock in one period - such as a drought - may well have
adverse repercussions on output in the years that follow, resulting in a
succession of negative error terms. Hence we may expect that the error
term in the model:
(11.9)
will probably not satisfy the assumption that E(E 1E1_5 ) O. Figure 11.4
shows the residual plot from estimation of equation (11.9) using a hypo-
thetical data set (given in Table 11.1); there are a number of fairly long
runs so the plot suggests that there may be a problem of autocorrelation
(a more formal test will be made: p. 384). There are procedures for

'¡¡j
o
::J
"O
·¡¡;
Q)
a:
-2

-4

-6

-8
1960 1962 1964 1966 1968 1970 1972 1974 1976 1978 1980 1982 1984 1986 1988

Figure 11.4 Residuals from crop production function


372 Econometrics far developing countries
Table 11.1 Hypothetical data set for crop production function
Output Price index Fertilizer input Rainfall
(Q) (P) (F) (R)
1960 n.a. 100.0 100.0 184.2
1961 40.4 106.0 99.4 155.3
1962 36.4 108.1 100.8 107.3
1963 35.4 110.3 102.1 110.1
1964 37.9 110.1 102.9 169.3
1965 34.8 108.6 103.1 81.9
1966 27.9 103.8 104.2 22.0
1967 29.8 109.5 104.6 31.5
1968 34.7 102.6 105.6 198.2
1969 38.4 101.1 106.8 147.5
1970 33.6 100.9 106.7 76.5
1971 33.6 104.7 108.3 90.2
1972 32.2 107.3 108.6 54.3
1973 35.3 103.0 110.4 178.3
1974 39.4 116.4 111.2 70.6
1975 30.6 112.7 111.1 53.0
1976 30.5 108.0 110.7 74.4
1977 33.7 103.2 110.5 136.4
1978 35.8 101.0 112.0 131.1
1979 36.0 103.6 111.6 109.2
1980 37.0 109.6 113.2 112.8
1981 30.7 105.2 114.1 43.0
1982 28.0 98.7 114.8 53.5
1983 28.4 99.2 114.8 20.6
1984 27.6 94.8 114.4 56.4
1985 32.9 100.6 114.6 78.3
1986 37.1 104.5 114.5 151.8
1987 36.0 98.9 114.2 145.2
1988 36.6 101.8 115.5 112.3
1989 38.8 105.6 116.7 161.0
1990 37.1 108.7 118.0 94.5

obtaining valid t-statistics under these circumstances, which are discussed


below (pp. 387-90). But first we must see how autocorrelation can result
from model misspecification.

Exercise 11.1
Use the data given in Table 11.1 to regress output on the price index and
fertilizer input. Draw the residual plot and count the number of runs.

lncorrect functional form


The residuals we get are a product of the model we impose on the data.
Consider Figure 11.5, which shows the population of Belize for the period
Misspecification and autocorrelation 373

200
190

180
Cñ 170
o
o
s. 160
e
o
:m::J 150
c.
o
a... 140

130

120

11 o
100'---'----'---'~~~-'----'--'--'---'--'----'--'-~---'---'~~~-'----'--'--'---'---'

1970 1972 1974 1976 1978 1980 1982 1984 1986 1988 1990 1992

- - Actual --+-- Fitted

Figure 11.5 Population of Belize and fitted line from linear regression

1970-92 and the fitted line from the linear regression of population on
time. Clearly the attempt to fit a straight line to these data is inappro-
priate - and the result can be clearly seen that the residuals are all positive
at first, negative in the middle years, and then positive in the last years.
This very clear pattern in the residuals is also displayed in Figure 11.6.
This case is one in which the autocorrelation results from the incorrect
functional form: the linear regression is not the right one, and this fact
affects the residuals so as to induce a pattern of autocorrelation. The auto-
correlation is a product of the incorrect functional form and not a property
of the error in the true model.
From looking at the graph we might suspect that the true model is to
regress the log of population on time. 4 As may be seen from Table 11.2,
this data transformation <loes improve the results. Unfortunately, as Figure
11.6 shows, the residuals from such a specification are a bit lower but still
display marked autocorrelation; there are still only three runs and the
results don't pass the rule of thumb for spurious correlation (R 2 > DW).
What is going on here?
If we look at the data it is possible to detect that the rate of increase
in population is lower in the earlier period (up to about 1977) than in the
later, whereas so far we have assumed the slope coefficient to be constant
throughout. To allow for this break in the data we regressed logged popu-
lation on time and an intercept and slope dummy with the break point
374 Econometrics far developing countries
10

-2

-4 l_L__J~L_L_.J.._-'-...J..._-'---'--'---'--__¡_--'----'-__J.__.JL_L_.J.._-'-...J..._-'---'-_J_J

1970 1972 1974 1976 1978 1967 1980 1982 1984 1986 1990 1992

----- Linear --+- Lag --- Lag with dummy

Figure 11.6 Residuals from regressions of population of Belize on time

Table 11.2 Regression results for regression of population of Belize on a trend


Dependent Regressor R2 DW
variable Constant Trend Intercept Slope
dummy dummy
POP 10.85 3.71 0.98 0.12
(70.57) (33.07)
ln(POP) 4.73 0.024 0.99 0.16
(781.25) (54.71) o. 99ª
ln(POP) 4.77 0.015 -0.075 0.011 0.99 1.81
(2477.92) (34.01) (-17.03) (25.54)
Notes:
t-statistics in parentheses
ª Square of correlation coefficient between exponential of fitted values and actual values.

in 1977. The residuals from this regression are also shown in Figure 11.6.
There are now ten runs (compared to the previous three runs) and the
danger of having to reject the null hypothesis of no serial correlation is
considerably reduced.
A word of caution must be inserted here. The econometric interpreta-
tion of our results is that there is a structural break in the regression of
Misspecification and autocorrelation 375
Belize's population on time, with a higher growth rate (slope coefficient)
in the later years than the earlier. But in fact we will often find such
results with population data. Typically a country's population is enumer-
ated once every ten years in a census; population figures for non-census
years are estimates based on the most recent census and observed intra-
census trends. When the data from a new census become available there
will be a break in the intercept as the actual figure will not equal the
estimate made on the basis of the previous census, and a new popula-
tion growth rate will be used for future estimates as this figure is also
revised. The structural break in this case is therefore a product of the way
in which the data are produced rather than any sudden change in the
proclivity of the people of Belize. None the less, the example serves
to illustrate how an inappropriate functional form may result in residual
autocorrelation and, also, how adding a further variable to a regression
may remove autocorrelation. It is to this latter possibility that we now
turn our attention.

Omitted variables as a cause of autocorrelation


We saw in Chapter 6 that the exclusion of relevant variables will bias the
estimates of the coefficients of all variables included in the model (unless
they happen to be orthogonal to the excluded variable). The normal t-
tests and so on. cannot tell us if the model is misspecified on account of
omitted variables, since they are calculated on the assumption that the
estimated model is the correct one. However, the presence of autocorre-
lation can be symptomatic of a problem of omitted variables, in which
case the least squares estimates are of course biased (Chapter 6).
Let us return to the crop production function discussed above. The
residual plot was presented as an example of the traditional view of auto-
correlation that it is found because of a genuine relationship between the
error term in successive variables. But there may be alternative explana-
tions for the pattern in the residuals in Figure 11.4, one of which is the
possibility of an omitted variable. Suppose we had included an indicator
of weather conditions in the production function and suppose further that
the variable was serially correlated (i.e. a good year is more likely to
follow a good year than a bad one - as in the biblical seven years of
plenty followed by seven years of drought). Omission of this variable
causes the model to underestimate output in years of good rainfall
and overestimate in those with poor weather conditions - and since good
years are bunched together we would expect to observe runs of posi-
tive residuals.
To examine this possibility we re-estímate the crop production function
including an index of rainfall (R) in the current period and once lagged
amongst our regressors. The resulting residual plot is shown in Figure
11.7. There is far less of a pattern in these residuals than previously (the
376 Econometrics far developing countries
3

-1

1960 1962 1964 1966 1968 1970 1972 1974 1976 1978 1980 1982 1984 1986 1988

Figure 11. 7 Residuals from crop production function from expanded regression

Table 11.3 Results from crop production function regression


Regressor
e p F R R(-1) R2 DW
10.21 0.26 -0.03 0.12 0.96
(0.41) (1.74) (--0.22)
-8.07 0.22 0.09 0.05 0.04 0.93 1.78
(-1.07) (4.95) (2.24) (11.46) (8.84)
Source: Derived from hypothetical data set given in Table 11.1.

number of runs is increased from 11 to 14), suggesting that the auto-


correlation was in fact a symptom of omitted variable bias. (It will also
be observed that the significance of the included variables has greatly
increased - see Table 11.3 - which also indicates that there was a problem
of omitted variable bias.)

Autocorrelation in cross-section data


Many textbooks will tell you that the problem of autocorrelation is partic-
ular to time series and consequently it need not be checked for in cross-
section data. It is correct that there cannot, literally speaking, be serial
correlation of the error in cross-section data, since the data are not a
Misspecification and autocorrelation 377
series; that is, there is no natural ordering in the way there is for time-series
data. None the less, if the data are ordered then misspecification may
show up as autocorrelation. (Genuine autocorrelation in cross section
data may also occur and is sometimes called spatial correlation - for
example, tastes may show a common pattern in the same street or village.
But here we are concerned with serial correlation as a symptom of mis-
specification.)
Consider the linear regression of life expectancy on income per capita.
Figure 11.8 shows the residual plot, where the observations along the axis
are ordered alphabetically by country. There is no readily apparent
problem in this residual plot. But Figure 11.9 shows the same residuals,
but now with the data sorted by income per capita. The autocorrelation
is very clear. The cause of the autocorrelation is shown in Figure 11.10,
which is the scatter plot for these data, where we are fitting a straight line
to an obviously non-linear relationship. The result is that the fitted line
lies above most observations for low and high income countries but below
those for middle income countries. Another lesson here is that if there
is genuine correlation in cross-section data it may be removed simply
by reordering the data: but you should be wary of so doing too quickly,
as the autocorrelation is most likely the clue to sorne misspecification.
Reordering is not usually an option in time-series data, and certainly not
so if the equation includes any lags.
20

15 ,__ D
D D D
D i:ri D D ~ D
10 ~
D D D
D oº
D
D CtJ D D D D
5r
D i:ri D D D
cR:i
D ºº Do D Cb D Do
rn::J D~ D ~D D D D
D
"O
·¡¡;
o D
D D D ro
r1 r1
D D u
D
CD D
a: D
D
D D
-5 ,__ D D
D Do Do D oº
00 D
Do D
D
D D
-10 ~ D
D D DO D
D ºº
D D D
D
-15 ~ D
D
D
-20 ~

1 1 1 1 1 1 1 1 1 1
-25
11 21 31 41 51 61 71 81 91 101
Observation (alphabetical)

Figure 11.8 Residuals from regression of life expectancy on income per capita,
alphabetical ordering
378 Econometrics for developing countries
20

15 D
D D D
D D ¡§11.l D D
10 - D
D C0 ci:P D
oº D ºo
Cf:¡
D D D D Do
5 - D Do
D ¡jl D D D D ºoo
D
Cii
:::J D
ºº
D
D C2:J D ctfü D
"O
·¡¡;
o D D D D
D
~~
(J)
a: D D

-5 - D D
º¿:i D D º!f:i
D D D
D
Doo D D
-10 - D D
D D
rn51oo
D D D
D
-15 - D
D
D
-20 -
1 1 1 1 1 1 1 1 1 1 1
-25
19 96 109 63 29 23 64 43 62 50 52
Observation (income)

Figure 11.9 Residuals from regression of lite expectancy on income per capita,
income ordering

90

80

70
>.
o
e
~
(J)
c. 60
)(
(J)

2
:.:::¡
50

40

30
o 2 4 6 8 10 12 14 16 18 20
lncome per capita ('000)
Figure 11.10 Scatter plot and fitted line for regression of lite expectancy on
income per capita
Misspecification and autocorrelation 379
Exercise 11.2
Compile time-series data for the population of the country of your choice.
(The data for Belize are in data file BELPOP and for Peru in PERU.)
Regress both population and logged population on time and graph the
residuals in each case. How many runs are there in each case? Comment.
Can you respecify the equation to increase the number of runs?

Exercise 11.3
Repeat exercise 11.2 using time series data for the country of your choice
(or from data file PERU) for: (a) real manufacturing value added; (b)
infant mortality rate; and (c) terms of trade. Comment on your results.

Exercise 11.4
Using the results in Table 11.3, use an F-test to test the hypothesis that
the two rainfall variables may be excluded from the regression. In the
light of your results, comment on the apparent problem of autocorrela-
tion in the regression of output on the price index and fertilizer input.

Exercise 11.5
Using data given in data file SOCECON, regress the infant mortality rate
on income per capita. Plot the residuals with the observations ordered:
(a) alphabetically; and (b) by income per capita. Count the number of
runs in each case. Comment on your results.

11.4 DETECTING AUTOCORRELATION


The first check for autocorrelation is to look at the residual plot, which
is a vital step after any regression. Another visual device is the correlo-
gram which graphs the covariances calculated from the residuals
(standardised by the variance). Correlograms play an important part in
identifying the appropriate model to fit to an autocorrelated error term,
but this discussion is beyond the scope of this book and we explain them
only in passing. We then move to two formal tests for autocorrelation:
the runs tests and the Durbin-Watson statistic. Finally, we introduce tests
to be used when there is a lagged dependent variable.

Correlograms
The correlogram is the plot of the residual covariances standardised by
the residual variance. By plotting the theoretical correlograms from
different error generating processes we can learn to spot these processes
380 Econometrics for developing countries
when confronted with these plots calculated using the residuals from actual
regressions. Here we will first consider the correlograms produced by the
error term generated using the AR(l) process, presented in Figures
11.1-11.3 above, and then plot those for the residuals from the crop
production function data.
Figures 11.11-11.13 show the correlograms for the data generated with
the AR(l) process for p =O, p = 0.7 and p = -0.7. When p =O we expect
the O term (i.e. the ratio of the error variance to itself) to be unity, as it
must always be, and all the others to be zero. In practice, the other terms
are not zero, but as Figure 11.11 shows they are comparatively small.
The lack of a pattern in Figure 11.11 stands in stark contrast to that shown
in Figure 11.12 where, as we would expect from equation (11.4) above,
there is a reduction in the covariances as we take errors which are further
apart. Equation (11.4) suggests that when p < O then the covariances
should alternate between negative (for odd differences) and positive (for
even differences) - and this pattern can be clearly seen from Figure 11.13
(for p = -0.7).
Turning from errors generated by a known model, we now plot the resid-
ual correlogram from estimation of the crop production function. Figure
11.14 shows the correlogram for the residuals from the regression of out-
put on the price index and fertilizer input. There is not such a marked pat-
tern as in Figure 11.12, but the high covariance between the residual and

1.20

1.00

0.80

0.60

0.40

0.20

O.DO

-0.20
o 2 3 4 5 6

Figure 11.11 Correlogram for AR(l) process with p = O


Misspecification and autocorrelation 381
1.20

1.00

0.80

0.60

0.40

0.20

0.00
o 2 3 4 5 6

Figure 11.12 Correlogram for AR(l) process with p = 0.7

1.50 ~----------------------------..,

o 2 3 4 5 6

Figure 11.13 Correlogram for AR(l) process with p = -0.7


382 Econometrics for developing countries
1.50

1.00

0.50

0.00

o 2 3 4 5 6

Figure 11.14 Correlogram for crop production function (regression 1)

o 2 3 4 5 6

Figure 11.15 Correlogram for crop production function (regression 2)


Misspecification and autocorrelation 383
its first lag are suggestive of positive autocorrelation. But above it was
argued that this apparent autocorrelation is not in fact a property of the
error term, but rather a symptom of omitted variables. It was found that
current and lagged rainfall had been incorrectly excluded from the regres-
sion. Figure 11.15 shows the correlogram for the regression including the
rainfall variables. The problem of autocorrelation appears to have been
removed.

Exercise 11.6
Using the results from your estimation of population against time (exer-
cise 11.2) plot the correlogram for the regression of: (a) population on
time; (b) logged population on time; and (c) your improved specification.

Runs test
So far we have been counting the number of runs as an indication of the
presence of autocorrelation: if there is positive autocorrelation then there
will be rather fewer runs than we would expect from a series with no
autocorrelation. On the other hand, if there is negative autocorrelation
then there are more runs than with no autocorrelation. But so far we have
not said how many runs are 'too many'. In fact it is possible to calculate
the expected number of runs from our residuals under the null hypoth-
esis of no autocorrelation and to construct a confidence interval around
this number. If the actual number of runs is less than the lower bound
we reject the null in favour of positive autocorrelation and if the number
of runs is above the upper bound we reject the null in favour of negative
autocorrelation.
It can be shown that the expected number of runs is:

E(R) = 2N1N2 +1 (11.10)


n
where N 1 is the number of positive residuals, N 2 the number of negative
residuals, R the total number of runs, and n the number of observations
(so n = N1 + N 2 ). The expected number of runs has an asymptotically
normal distribution with variance:

2 2N1N 2 (2N 1N 2 - n)
sR = nz(n - 1) (11.11)

Hence the 95 per cent confidence interval is given by:


E(R) - 1.96sR,;:;; R ,;:;; E(R) + l.96sR (11.12)
We accept the null hypothesis of no autocorrelation if the observed
number of runs falls within this confidence interval. The runs test table
384 Econometrics far developing countries
on pp. 474--5 gives the confidence intervals for a range of sample sizes. lf
the required values of N 1 and N 2 are not given in the table then equations
(11.10)-(11.12) should be used.
As an illustration we shall apply the runs test to two of the examples
given above: the crop production function and the regression of popula-
tion on a trend. The residuals from the regression of crop output on the
price index and fertilizer input have the following data: N 1 = 16, N 2 = 14
and R = 11 (these figures may all be read from Figure 11.4). Using N 1
and N 2 the 95 per cent confidence interval can be read from the runs test
table as 10-22. Since the observed number of runs falls within this range
we accept the null hypothesis of no autocorrelation. 5 With the crop
production function including the rainfall variables, N1 and N 2 are again
16 and 14 respectively (this fact is a coincidence, it need not be the case)
and R = 14. The observed number of runs is clase to the centre of the
confidence interval so we accept the null hypothesis of no autocorrelation
at the 5 per cent level.
For the case of Belize's population the simple log regression has N 1 =
14, N 2 = 9 and R = 3, compared to a confidence interval of 7-17 so that
we reject the null hypothesis of no autocorrelation. (Since there are fewer
runs than expected this is evidence of positive autocorrelation.) When the
dummies are added these values become N 1 = 14, N 2 = 9 and R = 10. The
confidence interval is 7-17 so that we accept the null hypothesis of no
autocorrelation.

Exercise 11.7
Use your population regression results to perform the runs test for auto-
correlation for the simple log regression and your preferred specification.

Exercise 11.8
Carry out the runs test on the errors shown in Figures 11.1, 11.2 and 11.3
and comment on your findings.

Durbin-Watson (DW) statistic


The DW statistic (d) is the most commonly reported test for autocorre-
lation, and is defined as:
T
L (et - e1-1)2
d = ~t=~2~T~-- (11.13)
L e;
t=l
Misspecification and autocorrelation 385
By multiplying out the top of this definition we get:

(11.14)

Making the following approximations:


T T T
L e1~1 L e7 "" L e7
t=2
""
t=2 t=l
(11.15)

it follows that:
/\ /\ Le1 e
d"" 2(1 - p) where p = t1 · (11.16)
L e1-1

where p is the slope coefficient from regressing the residuals on their first
lag. The Durbin-Watson statistic has thus been shown to be related to
the estimation of the AR(l) model, and it is indeed a test for first-order
correlation - that is, a non-zero covariance between the error and its first
lag. Alternative specifications of the test are required to test for higher
orders of autocorrelation, but first arder is what we meet most commonly
in practice. 6
From equation (11.16) we can see that if there is no autocorrelation
(so that p = O) then d = 2. As p tends to one (positive autocorrelation)
d tends to zero, and as p tends to -1 (negative autocorrelation) d tends
to 4. The Durbin-Watson statistic thus falls in the range O to 4, with a
value of 2 indicating no autocorrelation; values significantly less than 2
mean we reject the null of no autocorrelation in favour of positive auto-
correlation and those significantly above 2 lead us to reject the null in
favour of negative autocorrelation.
The critica! values for the DW test depend not only on sample size and
the number of regressors, but also the particular values of the regressors.
Consequently it is not possible to give a single critical value; instead two
values are given: an upper and a lower bound. The use of these bounds
is shown in Figure 11.16.
If the calculated value of d is less than the lower boundary (dL)
then reject the null hypothesis of no autocorrelation in favour of the
positive autocorrelation. If it líes above 4 - dL then reject the null in
favour of negative autocorrelation. (Since we more commonly find positive

Reject Ha Accept Ha Reject Ha

o du 2 4-du 4

Figure 11.16 Interpretation of the DW statistic


386 Econometrics far developing countries
autocorrelation the tables give only the values below 2; if the calculated
d is above 2, then you should work out 4 - dL and 4 - du.) If the calcu-
lated value lies between du and 4 - du then accept the null hypothesis of
no autocorrelation. The problem is if the calculated d lies between dL and
du (or 4 - du and 4 - dL); this is the 'zone of indecision'. To be on the
safe side we should reject the null, but we may decide not to if the value
is clase to du and we have other evidence (e.g. the runs test) against auto-
correlation.
In the case of the crop production function, the calculated DW is 0.96
far the regression of output on prices and fertilizer input. There are 30
observations so with three regressors (including the constant) as dL = 1.28
and du = 1.57. The calculated value is way below the lower bound crit-
ica! value (1.28) so we reject the null hypothesis of no autocorrelation in
favour of positive autocorrelation. But when we include the rainfall vari-
ables the calculated DW is now 1.78. Since the degrees of freedom are
reduced by the inclusion of the two extra regressors (R and R_1), the crit-
ica! values are changed, becoming dL = 1.14 and du = 1.74. The calculated
DW is closer to 2 than is du, so we may conclude that it is insignificantly
different from 2, reconfirming our argument that the autocorrelation was
a product of model misspecification.

Testing for autocorrelation with a lagged dependent variable


We saw above (section 11.2) that when one of the regressors is the lagged
dependent variable then OLS estimation is not only inefficient but also
biased. Hence detection of autocorrelation in this case is particularly
important. Unfortunately, the inclusion of a lagged dependent variable
also biases the DW statistic towards 2. Since 2 is the value of the statistic
in the absence of autocorrelation, this bias may lead us to accept the null
when in fact autocorrelation is present. Therefore the DW statistic may
not be used when the regressors include the lagged dependent variable
(but the runs test may). Durbin proposed another test, Durbin's h for
these circumstances, which is given by:

h -(1 -~) J
2
n
1 - n(Var (b1)]
(11.17)

where Var(b1) is the square of the standard error of the coefficient on the
lagged dependent variable and n is the number of observations. The test
may not be used if n·Var(b¡) is greater than one.

The DW statistic as a test for a unit root


The patterns displayed by a positively autocorrelated residual are the
same as those displayed by a non-stationary series around its mean; i.e.
Misspecification and autocorrelation 387
a stationary series would jump randomly around its mean value, whereas
a non-stationary one will stay away from it on apparent trends. If we
regress a variable on a constant then the estimated intercept from this
regression is the variable's mean, and the residuals thus describe the vari-
able's variation around the mean. The DW statistic from this regression
is therefore a test for stationarity, using the usual critica! values.

11.5 WHAT TO DO ABOUT AUTOCORRELATION


The main message in this chapter has been that autocorrelation may be
a symptom of misspecification. The first response to finding a problem of
autocorrelation with your regression is to conduct further specification
searches. In time-series analysis you should always first check if there is
a problem of spurious regression. Detection of spurious regressions was
discussed in Chapter 10 and estimation under these circumstances, using
an error correction model, is presented in Chapter 12. But in other cases
(or in estimating the levels regression for an error correction model) there
may be a problem of omitted variables and/or incorrect functional form.

Dealing with 'genuine' autocorrelation


The traditional view of autocorrelation is that in the model:
Yt = 131 + l32Xt + Er (11.18)
the assumption that E( E1E1_,) = O is violated, but the other assumptions
hold, including that the model is correctly specified. Throughout this
chapter we have stressed that it is very likely that the presence of auto-
correlation is in fact a signal that the model is not correctly specified.
Further specification searches should thus follow before suppressing the
autocorrelation. Suppose you are convinced that the model is the correct
one but the autocorrelation remains. Then under these circumstances you
may turn to the Cochrane-Orcutt procedure for producing estimates with
valid t-statistics.
The Cochrane-Orcutt procedure is the BLUE with an autocorrelated
error,7 and is as follows. Suppose that our model has an error which
follows an AR(l) process:
Yt = 131 + l32Xr + PE1_ 1 + v1 v1 - e(O,a2) (11.19)
If we lag equation (11.19) by one period, multiply through by p and
subtract the result from equation (11.19) we get:
(1 - p)Y1 = (1 - p) 13 1 + (1 - p) 13 2X 1 + v 1 (11.20)
So if we define new variables Y* =Y 1
- p Y1_1 and X/ =X 1
- pX1_1 to
estímate the regression:
(11.21)
388 Econometrics for developing countries
OLS is now efficient and 132 * = 132 • (Though the intercept must be worked
out as 13 1 = 13 1*/(1 - p).) 8
Calculation of the transformed variables requires a value of p, which
may be estimated from the residuals from the original regression or, more
simply, by applying the approximation d = 2(1 - p), so that p = 1 - d/2.
If the residuals from estimating equation (11.21) still show auto-
correlation the procedure may be repeated (i.e. estimate p from the new
residuals and define Y** and X** and re-estimate, etc.) until there is no
longer any autocorrelation (this is referred to as the Cochrane-Orcutt
iterative procedure ).
The differencing procedure used to calculate Y* and X* reduces the
sample size by one, since the first observation is lost. (If the first observa-
tion is Y1, this may be used to calculate Y 2*, but it is not possible to
apply the C-0 correction to calculate Y1 *.) The loss of this observation
is particularly serious in small samples (though we must also warn that
the Cochrane-Orcutt procedure is really only appropriate in larger
samples). The Prais-Winsten transformation is a formula for estimating
values of these lost initial observations; specifically: Y1* = (1 - p2) 112 Y1 and
X 1* = (1 - p2) 112X 1• Using these estimates allows us to preserve the sample
size. 9

An example: the crop production function


Regression of crop output on the price index and fertilizer input (reported
in Table 11.2) was found to be badly autocorrelated: the DW statistic
was 0.96 compared to a critica! value of dL of 1.28. We found that the
autocorrelation arose from a problem of omitted variable bias. But
for illustrative purposes we shall see how the autocorrelation may be
removed using the C-0 correction. To do this we carry out the following
steps:
1 The estimated equation with OLS gives DW = 0.958; thus p = 1 - d/2
= 0.521.
2 Calculate Q1* = Q1 - 0.521 Q1_ 1, and similarly for P* and F* for observa-
tions 1962 to 1990. The results are shown in Table 11.4.
3 Apply the Prais-Winsten transformation to get Q>IJ. 961 = (1 - 0.561 2) 112
·01961 , and similarly for the 1961 values of P* and F*. (Although we
do have 1960 values for P and F, though not Q, and so could
apply the C-0 procedure to the 1960 observations, the fact that we
use the Prais-Winsten transformation for one variable means that we
must also use it for the others.) The resulting values are shown in
Table 11.4.
4 Estimate:
(11.22)
Misspecification and autocorrelation 389
Table 11.4 Application of Cochrane-Orcutt correction to crop production
function data
Q Q* p P* F F*
1961 40.4 34.5 106.0 90.5 99.4 84.8
1962 36.4 15.4 108.1 52.9 100.8 49.0
1963 35.4 16.5 110.3 54.0 102.1 49.6
1964 37.9 19.4 110.1 52.6 102.9 49.7
1965 34.8 15.1 108.6 51.2 103.1 49.5
1966 27.9 9.8 103.8 47.2 104.2 50.5
1967 29.8 15.3 109.5 55.4 104.6 50.3
1968 34.7 19.1 102.6 45.5 105.6 51.1
1969 38.4 20.3 101.1 47.6 106.8 51.7
1970 33.6 13.6 100.9 48.2 106.7 51.1
1971 33.6 16.1 104.7 52.2 108.3 52.7
1972 32.2 14.7 107.3 52.7 108.6 52.2
1973 35.3 18.5 103.0 47.1 110.4 53.8
1974 39.4 21.1 116.4 62.8 111.2 53.7
1975 30.6 10.1 112.7 52.0 111.l 53.1
1976 30.5 14.5 108.0 49.3 110.7 52.8
1977 33.7 17.8 103.2 46.9 110.5 52.8
1978 35.8 18.2 101.0 47.2 112.0 54.4
1979 36.0 17.3 103.6 51.0 111.6 53.2
1980 37.0 18.3 109.6 55.6 113.2 55.0
1981 30.7 11.4 105.2 48.1 114.1 55.1
1982 28.0 12.0 98.7 43.9 114.8 55.4
1983 28.4 13.8 99.2 47.7 114.8 55.0
1984 27.6 12.8 94.8 43.2 114.4 54.7
1985 32.9 18.5 100.6 51.2 114.6 54.9
1986 37.1 20.0 104.5 52.1 114.5 54.8
1987 36.0 16.7 98.9 44.5 114.2 54.6
1988 36.6 17.9 101.8 50.2 115.5 55.9
1989 38.8 19.7 105.6 52.6 116.7 56.5
1990 37.1 16.9 108.7 53.6 118.0 57.2

Table 11.5 Regression results with Cochrane-Orcutt procedure


Constant p F R2 DW
OLS Coefficient 10.21 0.26 -0.03 0.12 0.96
(t-stat) (0.41) (1.74) (-0.20)
C-0 procedure Coefficient -9.57 0.28 0.22 0.62 1.50
(t-stat) (-2.02) (2.63) (1.58)

The regression results are given in Table 11.5 (which repeats also those
for OLS estimation). Calculate the estímate of the intercept b 1 = b 1*/(l
- p), which equals -19.98.
5 Comparing the two regressions, we see that the DW statistic is now
1.50. This value falls towards the upper end of the zone of indecision,
so the evidence for autocorrelation is much weaker than in the OLS
390 Econometrics f or developing countries
regression, though it may be thought worthwhile to repeat the pro-
cedure (using a new p of 0.25, calculated from the new DW).
Comparison of the slope coefficients from the two regressions shows
price to be relatively unaffected. With C-0 estimation, the fertiliser vari-
able produces the expected positive sign, though it remains insignificant.
The unexpected insignificance of fertiliser is a further indication that we
should have treated the initial autocorrelation as a sign of misspecifica-
tion. In this case, the C-0 procedure has suppressed the symptom of
misspecification, but cannot provide the cure - which is to include the
omitted variables.

11.6 SUMMARY OF MAIN POINTS

1 Autocorrelation is the violation of the classical assumption that the


error terms from different observations are unrelated to each other,
i.e. E( E;E) = O.
2 OLS with autocorrelation estimators is inefficient. Moreover, the
residual variance is not an unbiased estimator of the error variance so
that the t-statistics used for hypothesis testing are invalid.
3 If one of the regressors is a lagged dependent variable then OLS esti-
mates are also biased if the error is autocorrelated.
4 Traditional econometric practice interprets autocorrelation as a
problem of the model's error term. In fact the residuals may well
display a pattern of autocorrelation as a result of model misspecifica-
tion - either of omitted variables or incorrect functional form.
5 Since cross-section data may be reordered, any problem of 'autocor-
relation' may readily be removed by reordering the data. None the
less, if the data are ordered the presence of autocorrelation may also
be symptomatic of misspecification.
6 Various tests are available for autocorrelation. Visual methods include
the residual plot and correlogram. We also presented the runs test and
Durbin-Watson statistic. The DW may not be used if the model
includes the lagged dependent variable; in this case Durbin's h should
be used instead.
7 The first response to detecting a problem of autocorrelation in regres-
sion results should be to carry out more specification searches.
8 If you believe the model to be correctly specified and there is auto-
correlation, then the Cochrane-Orcutt procedure may be used to
obtain efficient estimates.
Misspecification and autocorrelation 391
ADDITIONAL EXERCISE

Exercise 11.9
Using the Sri Lankan macroeconomic data set (SRINA), perform the
simple regression of IP on Ig and plot the residuals. Use both runs and
DW tests to check for autocorrelation. Add variables to the equation to
improve the model specification (see Chapter 6 where this data set was
used previously) and repeat the tests for autocorrelation. Use the
Cochrane-Orcutt estimation procedure if you feel it is appropriate.
Comment on your findings.

APPENDIX 11.1: DERIVATION OF VARIANCE AND


COVARIANCE FOR AR(l) MODEL
The AR(l) process is:
(A.11.1)
Repeated substitution gives:
et = vt + p(pet-2 + vt_z)

= ~p;v (A.11.2)
~ t-1
i=O
Using equation (All.2) we may get:
=
e 1 e 1_ 8 _t-s~2i2
-p ~ p v- 1_ 8_¡ + cross pro ducts (A.11.3)
i=O
Therefore:

E(et et-s) = pt-s ~ p2; E(vt-s-i)


i=O
1-S 2
p ªV (A.11.4)
1 - p2

since the expected value of all the cross products is zero (as v 1 is not seri-
ally correlated). From which it follows that:
2
2)
E (e 1 = ª• 2
=
ªV
l _ p2
(A.11.5)
392 Econometrics for developing countries
NOTES
1 The data are generated using the Lotus @RAND command. These variables
follow a rectangular distribution, but are made approximately normal by aver-
aging over 20 such random numbers. (By the central limit theorem the resulting
numbers are approximately normal.)
2 Equation (11.3) is just the special case of equation (11.4) in which s =O.
3 The generalised least squares estimator (GLS) is BLUE, but further discus-
sion is beyond the scope of this text.
4 Such a specification is also desirable because of the interpretation of the slope
coefficient as the population growth rate.
5 However, the sample size is not that large and the value near the lower end
of the interval so we need to be cautious, perhaps by seeking verification from
another test. As we see below, the Durbin-Watson statistic suggests that these
residuals do show autocorrelation.
6 The exception worth noting is that quarterly time series data may well have
fourth-order autocorrelation.
7 And hence is equivalent to GLS, see note 3.
8 This fact means that we cannot readily use the t-statistic to test the signifi-
cance of the intercept when applying the Cochrane-Orcutt procedure. The
appropriate test is beyond the scope of this book.
9 The C-0 procedure is not equivalent to GLS, and therefore not BLUE, unless
the Prais-Winsten transformation is applied.
12 Cointegration and the error
correction model

12.1 INTRODUCTION
Thus far in Part IV we have discussed problems encountered in time-
series analysis: the danger of spurious regression in Chapter 10, and the
problem of autocorrelation more generally in Chapter 11. We have empha-
sised that regression with non-stationary series is generally biased and
inconsistent. Transformations to stationarity, notably differencing, create
their own problems. In this chapter we will present valid procedures for
obtaining regression estimates with non-stationary series. Least square
estimates can be used if two non-stationary series are cointegrated, a
concept we explain in section 12.2. The test for cointegration, of which
examples are given in section 12.3, is to test whether the residuals from
the levels regression are stationary. If these residuals are stationary then
the series are cointegrated. The levels regression will then provide consis-
tent estimates of the long-run relationship. The full dynamic model is
estimated as an error correction model, which is presented in section 12.4.
Section 12.5 concludes.

12.2 WHAT IS COINTEGRATION?


In Chapter 10 we saw how the regression of one random walk on another,
independently generated, random walk could yield significant results, even
though we know for a fact there is no relationship between the two vari-
ables. This result indicates the danger of spurious regressions when using
time-series data: if two series are non-stationary, then an observed rela-
tionship between them may be spurious. We also saw how such spurious
relationships collapse when the model is estimated in differences. How
can we spot if a regression is spurious or not? The differences test,
presented in section 10.5, is one method. More common is to test for the
cointegration of a set of variables.
In general, the sum or difference of I(l) variables will also be I(l). Two
I(l) variables, X and Y, are said to be cointegrated if there is sorne linear
combination, Y - bX, which is I(0). 1 This linear combination may be found
394 Econometrics for developing countries
by the OLS regression of Y on X, so that the residuals from this regres-
sion are equal to Y - bX. 2 Thus, simply put, variables are cointegrated
with one another if the residuals from the levels regression are stationary.
If variables are cointegrated then the levels regression is not spurious:
the OLS estimates are consistent (see Box 13.1 in the next chapter for a
discussion of consistency). Indeed, they are 'super-consistent', meaning
that the probability distribution converges on the population value more
rapidly than with I(O) variables. However, the usual t-tests do not apply.
Although the levels regression is consistent, it is usual to proceed to esti-
mation of the error correction model (ECM), which contains information
on both the long- and short-run relationship between the variables. We
discuss estimation of the ECM in section 12.4, but first we discuss the
intuition behind cointegration and how to test for it.

Cointegration: an intuitive approach


Why are stationary residuals indication of a non-spurious relationship?
We generated three variables, X, Yl and Y2 as follows:
(12.1)

(12.2)

Y2 = -2 + 0.8Xt + E2,t (12.3)


That is, X and Yl are both random walks without drift, and Y2 is a func-
tion of X. Regression of Y1 on X and of Y2 on X yields:
Yl 1 = -0.24 + 0.55X1 R 2 = 0.35
(-0.25) (5.18) DW = 0.35 (12.4)

Y2t = -1.80 + 0.78X1 R 2 = 0.81


(-3.62) (14.46) DW = 1.90 (12.5)
Both sets of regression results seem reasonable. But we know for a fact
that the results in equation (12.4) are spurious, whereas those in equa-
tion (12.5) are a good estimation of the actual data generation process.
If we compare Figures 12.1 and 12.2, which show the actual and fitted
values from the spurious and true models, we can see (a) why spurious
regressions can produce good R2s; and (b) how a spurious model is distin-
guished from a true one.
Equation (12.4) shows a reasonable R2 because X and Yl share the
same rough trends. In the first half of the sample period the match between
X and Yl is not that clase, because neither shows too strong a movement.
But from around period 24 both series embark on a downward drift. This
downward drift is reversed for Yl after period 40 and for X a few periods
later. What of the fitted value of Yl? Since Yl is regressed on the single
Cointegration and the error correction model 395

o 4 8 12 16 20 24 28 32 36 40 44 48
oX + Y1 o Fitted Y1

Figure 12.1 Actual and fitted values of spurious regression model


11 .-~~~~~~~~~~~~~~~~~~~~~~~~~~~---,

10

o 4 8 12 16 20 24 28 32 36 40 44 48
oX + Y2 o Fitted Y2

Figure 12.2 Actual and fitted values of true regression model


396 Econometrics far developing countries
variable, X, the movements in fitted Yl will precisely match the move-
ment of X, with the intercept term shifting the whole series down so that
it lies over the actual Yl series. Since, broadly speaking, X and Yl move
roughly the same, by following the trends in X, then fitted Yl follows Yl
quite closely, resulting in a reasonable R 2 •
Recall that the R 2 is percentage of the total sum of squares 'explained'
by the regression model, where the total sum of the squares measures the
variance around the mean. As they are non-stationary, both X and Yl
deviate from their mean as they have wandered away from it by following
a random walk. It is these common general drifts that cause X and Yl to
deviate from their means by a similar amount, rather than any causal link
between the two variables.
To see this, look more closely at Figure 12.1 and compare it to Figure
12.2. If there is a true relationship between the two variables, then the
dependent variable should, in general, move in the same direction as the
regressor (i.e. the relationship should also hold for differences) - though
this may not be so if there is a large error term in either the current or
preceding period. By contrast, if there is no relationship between the two
variables then moving in the same direction is a matter of chance. A
random walk has an equal probability of going up or down, and so the
probability that two random walks will move in the same direction is one-
half. Analysis of X and Yl shows the two variables to move in the same
direction in 27 of the 50 cases: that is, as expected, approximately half.
By contrast, X and Y2 move in the same direction in 37 of the 50 cases.
Herein lies the essential difference between a true relationship and a
spurious one. A random walk follows an apparent trend, and the resulting
deviation from the series' mean explains the largest part of its variation.
Hence another random walk, which also follows an apparent trend, will
get similarly large deviations from its mean, and so one series may 'explain'
the general movements in the other quite well. But whilst the general move-
ments are explained, the actual period-on-period movements will not,
except by chance, match. Hence if we look at a graph in levels, the fitted
values of Yl will be at the same general level as those of Yl, but if we look
more closely we see that fitted Yl moves in the same direction as Yl only
about half of the time. This fact is why a spurious regression collapses
when the model is estimated in differences. The contrast with Figure 12.2,
in which the direction of movement matches in 37 cases, is clear.
What are the implications for the residuals? The residual plots from
equations (12.4) and (12.5) are shown in Figures 12.3 and 12.4 respec-
tively. The former appears to be clearly autocorrelated, whilst the latter
is not. Autocorrelation arises in the spurious regression because the fitted
values move only in roughly the same direction as the actual values, but
do not match it precisely. To see how this fact results in autocorrelation,
consider the other case of the true model. The deterministic part of the
model will give fitted values of Y2 that almost exactly match the actual
Cointegration and the error correction model 397
series Y2 (not exactly, as the regression estimates are close, but not equal,
to the true model values ). Whether the residual is positive or negative
then depends if the error is positive (taking Y2 above the fitted value) or
negative (taking it below) - and, most importantly, which of these two
events occurs is unrelated to what happened in the previous period. Hence
the residuals will show no evidence of autocorrelation. In Figure 12.2 the
fitted line jumps either side of the actual values in a random manner, as
shown by the quite jagged residual plot.
This argument does not apply at all to the case of a spurious regres-
sion. Suppose the actual and fitted values of Yl coincide at a particular
point in time - as they do in period 28. After period 28 both series embark
on a downward drift for four periods - both e1 and e2 are negative (except
e2 in period 30). But the drift in Yl is stronger than that in X, so the
fitted values of Yl lie above those of actual Yl. Moreover, as a random
walk remembers where it is, even when the X series is subject to a series
of positive shocks, the fitted values of Yl remain above Yl for sorne time.
Hence, as clearly shown in the residual plot, there is a long run of nega-
tive residuals from period 28 right through to period 42. Fitted Y1 follows
X, which is not related to the movements in actual Yl, so the fitted values
can remain too high or too low for quite lengthy periods, resulting in the
pattern of autocorrelation shown in the residual plot from estimating equa-
tion (12.4). Thus we see why autocorrelation may well be the symptom
of a spurious regression.

1.1 ~~~~~~~~~~~~~~~~~~~~~~~~~~~

1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.01--~-+-~~~~~~--+~++---i'--+---~~~~~--i~-+-~~~~

-0.1
-0.2
-0.3
-0.4
-0.5
-0.6
-0.7
-0.8
-0.9
-1.0
-1.1 l..--'-~--'-~--'-~-'-~'----'~-'-~-'-~-'-~-'--~+-------''-----'-~~~

o 4 8 12 16 20 24 28 32 36 40 44 48

Figure 12.3 Residuals from spurious regression


398 Econometrics for developing countries
0.1~~~~~~~~~~~~~~~~~~~~~~~~~~~

0.6
0.5
0.4
0.3
0.2
0.1
O.O 1----+--i-+--+~-+--+-~+-+-_,_.'-+-+-+--++-++~--+--+-~-+~--+~1---<.-+-~~

-0.1
-0.2
-0.3
-0.4
-0.5
-0.6~~~~~~~~~~~~~~~~~~~~~~~-+~~~

o 4 8 12 16 20 24 28 32 36 40 44 48

Figure 12.4 Residuals from estimation of true model

The most common test for autocorrelation is the DW statistic, which in


its usual form is based on modelling the error as an AR(l) process.
Compare the residual plots in Figure 12.3 and 12.4 with the plots of the
AR(l) process with different parameter values given in Figures 10.4-10.10.
The residuals from estimation of the true model match very closely the plot
shown as Figure 10.4: that is, the AR(l) model with a zero autoregressive
coefficient. You might think that Figure 12.4 is also similar to Figure 10.5 -
which shows an AR(l) process with an autoregressive coefficient of 0.5. In
either case, the matching plots suggest that the residuals shown in Figure
12.4 are stationary. The residuals from the spurious regression seem far
more like a random walk (Figure 10.6) - displaying the patterns that come
from a series remembering where it is. So we find that the residuals from
the estimation of a true modelare stationary, whereas those from a spuri-
ous regression are not. Therein lies the basis of the cointegration test.

Exercise 12.1
Using either the data in data file PERU or data for the country or coun-
tries of your choice (be sure to use real macroeconomic data, not nominal),
plot (a) the dependent variable, independent variable and fitted values;
(b) the residual plot; and (c) the dependent variable and independent
variable both expressed in differences, for the following relationships:
Cointegration and the error correctíon model 399
1 consumption as a function of income;
2 imports as a function of income;
3 the infant mortality rate as a function of income;
4 real growth as a function of export growth.
In the cases where a significant relationship is found, which of these rela-
tionships do you believe to be spurious?

12.3 TESTING FOR COINTEGRATION


There are two approaches to testing for cointegration. The first is an
analysis of the stationarity of the residuals from the levels regression, and
it is this approach, which is valid for bivariate analysis, that we pursue
here. But in multivariate analysis there can be k - 1 cointegrating vectors
(where k is the number of regressors). The Johansen method is a test for
cointegration which also determines the number of vectors. The latter
technique is not dealt with here, though the reader is referred to treat-
ments of the topic below.
lt was shown in section 12.2 that two variables are cointegrated if the
residuals from the levels regression are stationary. Hence the first step in
testing for cointegration is to estimate the levels equation:

(12.6)
This equation is called the cointegrating regression. We analyse here only
the bivariate case. The analysis changes if more variables need be included
on the right-hand side, but the technique is beyond the scope of this book.
If the residuals from the cointegrating regression are stationary then the
variables are said to be cointegrated. We know that the residuals will have
a zero mean and no trend by construction, so we do not need to apply
the full decision tree presented in Chapter 10. Rather we can proceed
directly to the augmented Dickey-Fuller test without a constant or a trend
or use the DW-based test. However, when testing for the stationarity of
residuals the critica! values used in Chapter 10 are no longer valid. The
correct ones to use, given below in the statistical tables on p. 480, are
slightly larger.

Sorne examples

The response of Pakistani exports to real devaluation


As a first example, consider the export supply function for Pakistan (data
file PAKEXP). Figure 12.5 shows the logged value of real exports. The
figure also shows the logged real exchange rate index (RER); that is a
weighted average of the index with Pakistan's main trading partners which
also allows for differential price movements. The exchange rate is defined
400 Econometrics far developing countries
as local currency per units of foreign currency, so that an increase in the
RER is a devaluation. There was a sharp real devaluation in the early
1970s, which was subsequently eaten away by higher infiation. Since the
mid-1980s the government has pursued an adjustment programme which
has resulted in a sustained real exchange rate depreciation. It appears
from the figure that exports have followed these movements in the RER
quite closely, suggesting that there may be a causal link between the two.
This view is supported by the scatter plot (Figure 12.6) which shows a
close relationship between them, with only one outlier (1972) and no
readily discernible infiuential points.
To analyse this relationship further we regressed log exports on the
logged real exchange rate. OLS estimation yields (t-statistics in paren-
theses ):
ln(X) = -2.06 + 0.91ln(RER) R 2 = 0.71
(-3.63) (7.89) DW = 0.82 (12.7)
We might be pleased with the result. The price elasticity of supply of
exports is 0.91 - a 1 per cent devaluation increases exports by 0.9 per
cent. The R 2 is good, though the DW statistic may be a bit of a worry.
In fact, more than being a bit of a worry, the low DW is a clue to the
fact that these results are wholly spurious.
Befare estimating equation (12.7), the first thing we should have done,
as with any time-series data, is to check the order of integration of the
variables. Table 12.1 reports the results of the relevant tests. These results

2.9 5.4

2.8 5.3
•'"'
•' ....... ,,,.,. , r
'' '' o
2.7 '' '' 5.2 ce
''' ''' ' ce
x
¿ 2.6 ,..............:''
(1)
a.
(/)
5.1 ii3
t:: ' ~
o 2.5 '' (1)
a.
>< '' 5.0 ><
()
Q)
2.4 '' ::r
e¡; '' O>

~
'' '' ::::l
4.9 ce
2.3 '' (1)
'O '' ''
''
ª
Q)
'' 4.8
''
Ol (1)
Ol '
o 2.2 ',' '
',, / ?
,,,
...J

2.1 ---' 4.7 JJ


m
' / ~
'' ,'' 4.6
2.0 ,___ ,
'
1.9 4.5
66 68 70 72 74 76 78 80 82 84 86 88 90 92

1 -LX ·-----LRER 1

Figure 12.5 Pakistaní exports and real exchange rate, 1966-92


Cointegration and the error correction model 401
3.00 ~---------------------~

+
+
2.75
+
+++
+
++ +
2.50 +
+
X + +
....J ++

2.25 + +
+
+ +
++
2.00 +

1.75 -t---------r-----.-----~-------1

4.50 4.75 5.00 5.25 5.50

LRER
Figure 12.6 Scatter plot of exports and RER

are laid out similarly to Table 10.3, where we first met the decision tree
for testing for stationarity (except that here we only require the top two
rows of the table). Both variables are found to be I(l). Thus there is a
danger that OLS estimates using these variables may be spurious. To
examine whether or not our results are indeed spurious we must examine
the stationarity of the residuals from equation (12.7).
Since the residuals must have zero mean and no time trend we may
use the augmented Dickey-Fuller test in which the change in the variable
is regressed upon its own lag and a lagged difference term (plus addi-
tional lagged differences if they are necessary to remove residual autocor-
relation). The null hypothesis that the variable has a unit root is the test
of significance of the lagged term, so the normal t-statistic is calculated.
We first regress the change in the residual on lagged residual and check the
DW (this is just the Dickey-Fuller test); if there is autocorrelation we add
the lagged difference and re-estimate. In fact the DW without the lag
is 2.03, so there is no need to proceed to the augmented version of the test.
The t-value in this equation is -2.54 (shown in Table 12.1). The critica!
402 Econometrics far developing countries
Table 12.1 Decision tree applied to exports, RER and residuals from cointe-
grating regression
Exports RER Residuals
RSSu 0.352 0.374
RSSRI 0.608 0.527
(F-stat) (5.33) (3.00)
t-stat -2.54
Result Random walk Random walk Random walk
Note: Exports and RER are mean deviations of logged values.

values for residuals are not the standard Dickey-Fuller ones but are slightly
larger. At the 5 per cent level the critica! value given by Engle and Granger
(1987) is -3.37 for the DF test (and-3.37 for the ADF); Phillips and Ouliaris
(1990) give a slightly lower value of-2.76 (though this value is derived from
much larger samples). 3 Using either value, the null hypothesis is accepted;
that is, the residuals have a unit root so the two series are not cointegrated.
We mentioned in Chapter 11 that an alternative test for non-stationarity
is to regress the series on a constant and look at the DW statistic. When
applying the DW test to a series the result tells whether the variable is sta-
tionary or not, but yields no additional information as to the appropriate
dynamic specification. But, as already stated, in the case of residuals we
know there is a zero mean and no time trend and we are, in any case, only
interested to know whether they are stationary or not. The cointegrating
regression DW ( CRDW) is simply the DW statistic from the levels regres-
sion; in this case the regression of ln(X) on ln(RER). Recall that DW = 2(1
- p). Hence the null hypothesis that p = 1 (i.e. there is a unit root) corre-
sponds to the null DW =O. Engle and Granger (1987) give the appropriate
critica! value at the 5 per cent level to test this hypothesis as being 0.386.
In the example given here the calculated value is 0.824. As the calculated
value is greater than the critica! value we should reject the null of a unit
root in the residuals and thus conclude that the series are cointegrated.
The two tests, ADF and CRDW, thus give different results. Engle and
Granger say that the ADF is the recommended test. The critica! values
of the DW are in fact very sensitive to parameter values. Hence whilst
CRDW may be a quick and easy test to perform, the ADF (which does
not take much longer anyway) is preferable. Hence we conclude that the
results reported in equation (12.7) are spurious. It would be wrong to
conclude on the basis of these results that exchange rate policy has been
the driving force behind Pakistan's recent export growth.

A consumption function far Costa Rica


Figure 12.7 plots real GDP and real consumption for Costa Rica for the
period 1963-93 (data file CRCON; fitted consumption is also shown, but
Cointegration and the error correction model 403
5.8
5.7
5.6
5.5
e 5.4
o 5.3
aE 5.2
::::i 5.1
en
e 5.0
o
(.) 4.9
"O
e 4.8
ro 4.7
Q)
E 4.6
o(.) 4.5
.S 4.4
"O
Q)
4.3
Cl 4.2
Cl
o
...J
4.1
4.0
3.9
3.8
3.7
3.6
1963 1966 1969 1972 1975 1978 1981 1984 1987 1990
o = Consumption + = Fitted Consumption <> = lncome
Figure 12. 7 Costa Rican consumption function: actual and fitted values

ignore this line for the moment). Even from just looking at this graph we
should at least suspect two things. First, regression of consumption on
income is going to yield a high R 2 (almost certainly in excess of 0.9).
Second, this seemingly good result will probably be spurious, as closer
examination shows that, although both series 'trend' upwards, the year-
on-year changes in consumption do not in fact match particularly well
with the changes in income. That the residuals will be autocorrelated is
already clear from this plot, especially for the 20 years to 1982, during
which there are only three runs.
Our suspicions should be further aroused by the scatter plot (Figure
12.8), from which the autocorrelation is shown by the pattern of points
around the fitted line. This indication of autocorrelation is supported by
the regression results, which are (t-statistic in parentheses ):
In(C) = -1.53 + 1.23ln(Y) R2 = 0.98
(-10.42) (42.15) DW = 0.36 (12.8)
Investigation reveals both logged consumption and income to be
random walks without drift. Hence, the estimated consumption function
is indeed spurious unless the residuals from the levels regression turn out
to be stationary. We see from equation (12.8) that the regression fails the
CRDW as the DW is less than the critical value of 0.38. This result is
confirmed if we carry out the augmented Dickey-Fuller test. As befare
the first stage is the Dickey-Fuller test, but the DW statistic from this
404 Econometrics far developing countries
5.5
5.4
5.3 o
5.2
5.1 o
5.0
e 4.9
o
"üi 4.8
o.
E 4.7
::J
en 4.6 o
e
o 4.5 o
u
"O o
Q) 4.4 o
Ol
Ol 4.3 o
o
_J
4.2
o
4.1
4.0
3.9 o
3.8 o
o
3.7
3.6
4.2 4.4 4.6 4.8 5.0 5.2 5.4 5.6
Logged income

Figure 12.8 Scatter plot for Costa Rican consumption against income

regression is 1.03, suggesting that we need to add sorne lagged differences


to the right hand side. Adding one such term increases the DW to 2.24
and the t-statistic from this equation to test the null of no cointegration
(i.e. non-stationary residuals) is -2.94. Although in absolute terms this
figures is above the critica! value of Phillips and Ouliaris it falls below
that of Engle and Granger so we are safest to accept the null hypothesis;
i.e. the variables are not cointegrated.
In preparing this chapter we estimated consumption functions using
annual data for Ecuador, Finland, France, Greece, Pakistan, Philippines,
Trinidad and Tobago, and the United Kingdom. All these regressions
yielded good fits which turned out to be spurious! This fact underlines
the importance of allowing for the effects of non-stationarity in time-series
analysis. Our failure to find a non-spurious relationship <loes not mean
that consumption is unrelated to income. Several reasons may explain our
result. One important one is likely to be that the real income measure
we had available was GDP: whereas, especially for developing countries,
it is more appropriate to use GNP (which includes net factor payments)
or national disposable income (which includes grant aid received from
aid donors and workers' remittances). Another limitation is that we have
been using annual data. Quarterly data are rarely available for devel-
oping countries (other than for money and prices). But accurate modeling
of consumption probably requires quarterly data, the true relationship
Cointegration and the error correction model 405
being drowned out by aggregation to annual figures. Or maybe the data
generation process reflects sorne more complicated relationship between
consumption and income, such as the permanent income hypothesis.

A cointegrating consumption function


Using the data for the period 1960-93, in data file COINCON, the
following consumption function was estimated (using logged data, t-statis-
tics in parentheses):
ln(C) = -0.11 + 0.97ln(Y) R 2 = 0.99
(-0.98) (42.62) DW = 1.68 (12.9)
Is this relationship spurious? A look at the actual and fitted values
(Figure 12.9) suggests that it may not be so. In Figure 12.9 we see that,
although the actual values fluctuate around the fitted ones, they do so
in an apparently random manner. This pattern may be contrasted with
that shown in Figure 12.7, in which there appears to be a more system-
atic variation between the actual and fitted values. (This comparison is
the same as that made between Figures 12.1 and 12.2, showing the spurious
and true regressions.)

_____ . ., .
5.50

5.25

---
5.00

4.75

4.50-+-...,.---,-,--,.--,-,-,-,--,--...,.---,-,---,---,-,-,-,--,--...,.---,-,---,---,-,-,-,--,--;-,-,---,---.--I
1960 1965 1970 1975 1980 1985 1990
- - LC ------LY - - - LCFIT 1

Figure 12.9 Fitted and actual values for consumption function


406 Econometrics for developing countries
The residuals from equation (12.8), do indeed turn out to be stationary.
The DW is greater than the critica! value for the CRDW test. The t-
statistic from the DF test (the ADF is unnecessary since the DW from
this regression is 2.02) is -4.64, thus decisively rejecting the null hypoth-
esis of non-stationary residuals. Hence the two series are cointegrated.
The levels regression may thus be interpreted as a consistent estímate of
the long-run relationship between the two variables, an interpretation we
pursue below in the context of the error correction model.
The coefficient between logged consumption and logged income has
been found to be approximately unity. Since the coefficient in a double
log function is the elasticity, we have found a unit elasticity of consump-
tion with respect to income; that is, a 1 per cent rise in consumption will,
in the long run, result in a 1 per cent rise in consumption. Thus, there is
a long-run stable value of the average propensity to consumer ( CJY). We
say that there is a long-run relationship, since the estimated equation <loes
not have a lag structure, and we may expect there to be an adjustment
period as economic variables adjust to their equilibrium values. The error
correction model captures both the long- and short-run relationships
between two variables.

A final word on cointegration


We have presented here cointegration analysis for two variables. Turning
to multivariate analysis there may be more than one cointegrating vector.
Unfortunately, tests for the number of such vectors (of which the Johansen
method is the most common) are beyond the scope of this book. The
interested reader is referred to Rolden and Perman (1994) and papers
in the volume edited by Rao (1994). A more comprehensive, but very
technical, reference is Hamilton (1994). An important difference between
bivariate and multivariate cointegration analysis is that in bivariate
analysis two variables must have the same arder of integration in arder
to be cointegrated. This condition need not hold for all the variables in
a multiple regression; the set of variables can be cointegrated even though
the arder of integration varíes.

Exercise 12.2
Use the cointegration test to determine if the relationships listed in exer-
cise 12.1 are spurious for your data.

12.4 THE ERROR CORRECTION MODEL (ECM)


Recall the problem we faced. We wish to estimate the relationship between
two variables - in this case consumption and income. Yet these two vari-
ables are non-stationary, and we know that regression of non-stationary
Cointegration and the error correction model 407
series can give spurious results. What are we to do? Since both variables
are 1(1) then, by definition, their differences will be stationary and we may
regress the change in consumption on the change in income. For sorne
years, the accepted wisdom with time series was therefore that it was best
to work with the variables expressed in differences. However, such a regres-
sion yields no information on the long-run equilibrium relationship
between consumption and income.
Why is this? To salve a dynamic equation for its equilibrium values we
set x* = x 1 = x1_ 1, etc. Using this technique all difference terms become
equal to zero - there is, after all, no change in equilibrium. If a model
contains only differences, then all the variables 'drop out' when we try to
salve for equilibrium values! While a model in differences gets round the
estimation problem we face, it does so at the cost of not being able to
discover the long-run relationship between the variables of interest.
Cointegration analysis allows us to avoid paying this price.
The error correction model combines long-run information with a short-
run adjustment mechanism. To do this, define an 'error correction term',
EC, which is nothing other than the residual from the levels regression.
The error correction term is then used to estimate the ECM:
(12.10)
Since both difference terms and the error correction term are all 1(0) we
may estimate equation (12.10) by OLS. (For an alternative view of the
ECM, see Box 12.1.)
Estimation using the consumption function data from the last section
yields (t-statistics in parentheses ):
A

tiLC1 = 0.88tiLY1 - 0.94ECt-l R2 = 0.52


(8.81) (-5.27) DW = 1.93 (12.11)
lt is a relief to arrive at an econometrically sound estimate of the rela-
tionship between consumption and income. But this is rather far removed
from the simple consumption function with which we are familiar. How
are we to interpret the result in equation (12.11)? How can we learn from
it the marginal and average propensities to consume?

Interpreting the error correction model


We shall discuss interpretation of the ECM in two stages. First, we shall
explore the model's dynamics with sorne simulations. Second, we shall
interpret the parameters of the consumption function estimated in equa-
tion (12.11).
There are two parts to the dynamics of the ECM: the impact effect and
the error correction process. Figure 12.10 shows an exogenous variable
X, and the endogenous variable Y. Y has been calculated by a determin-
istic ECM (i.e. with no stochastic element) in which ¡3 1 = O, ¡32 = 0.6 and
408 Econometrics far developing countries

Box 12.1 The error correction model as a restricted time


series model
The error correction model may be arrived at in two ways. The one
we are concerned with in the text is as a convenient and valid repre-
sentation of non-stationary series. But it may also be seen as a
specific restriction of a more general model.
The general model (called the autoregressive distributed lag,
ADL) is:
Yt = 131 + l32Xt + l33X1-1 + 134Y1-1 + et
Spanos (1986: 552) lists eight classes of model which may be derived
by various parameter restrictions on this model. The one of interest
to us is 132 + 133 + 13 4 = l.
First, subtract Y1_1 from both sides:
LlYt = 131 + l32Xt + 133Xt-l + (1-134) Yi-1
But the restriction implies 13 3 = 1 - 13 2 - 13 4. Therefore,
Ll yt = 131 + l32Xt + (1 - 132 - l34)XH + (1 - 134) yt-l
= 131 + l32LlXt +(1 - 134) (Xi-1 - Yt-1)
which is the error correction model.

13 3 = -0. 75. The long-run relationship is Y* = 0.9X*, so that EC = Y -


0.9X. Initially X is constant with a value of 10. Between periods 5 and 6
there is a step increase to 11. The ECM model says that the change in Y
in period t equals the change in X in period t plus a correction for the
discrepancy between Y and its equilibrium value in period t - l. In period
5 the model was in equilibrium, with Y at 9 ( = 0.9 x 10), so the error
correction term plays no part in the change in Y in period 6. The coeffi-
cient on LlX1 is 0.6, so that Y increases by 0.6 units ( = 0.6 x 1). From
period 7 onwards X is constant, so that LlX is zero, and this term plays
no part in the adjustment process.
Since the equilibrium relationship is Y = 0.9X, the adjustment in Y of
0.6 units in period 5 is insufficient to restare equilibrium. In period 6, X
is 11 but Y is only 9.6, rather than its equilibrium value of 9.9. There is
therefore an 'error' of -0.3 ( = 9.6 - 9.9.) Since Y is below its equilibrium
it needs to increase for the model to converge toward equilibrium. This
movement is assured by 13 3 < O. (Similarly if Y > Y* then 13 3 < O ensures
that Y declines toward equilibrium). As we have 13 3 = -0.75, three-quar-
ters of the discrepancy between Y and Y* is corrected in each period
(specifically, Ll Y7 = -0.75 x -0.3 = 0.225, so that Y7 = 9.825). Hence, as
shown in Figure 12.10, Y converges quite rapidly on to its equilibrium.
Cointegration and the error correction model 409
11.6
11.4
11.2
11.0
10.8
10.6
10.4
10.2
a.>
O>
o 10.0
_J 1

(
9.8
9.6
9.4
9.2
9.0
8.8
8.6
8.4
o 2 4 6 8 10 12 14 16 18 20

Time
oX +Y

Figure 12.10 Simulation of error correction model (rapid adjustment)

Figure 12.11 shows the same simulation, but now with ¡3 3 = -0.25, so
that only one-quarter of the adjustment process occurs in each period. It
now takes rather longer for equilibrium to be restored. If the impact effect
is larger than the long-run effect then the model simulations will demon-
strate 'overshooting'.
In these simulations we have ignored the intercept, ¡3 1. We should allow
the value of this coefficient to be determined by the data. If ¡3 1 is non-
zero, then ¡3 1 and ¡3 3 become involved in the equilibrium condition,
suggesting that the dependent variable is subject to sorne drift in addition
to the equilibrium relationship and opening a question as to what we mean
by equilibrium. Equilibrium as used in the context of the cointegrating
regression and the ECM means a statistically observed relationship
between the variables over the sample period. This concept does not
necessarily correspond to economic equilibrium.
Turning to the estimated values for the consumption function shown in
equation (12.11), we can first note that it was valid to exclude the inter-
cept, and that the estimates of ¡3 2 and ¡3 3 have the expected sign. The
model converges quickly to equilibrium, with over 90 per cent of the
discrepancy corrected in each period. What can we say about the marginal
and average propensities to consume?
410 Econometrics for developing countries
11.6
11.4
11.2
11.0
10.8
10.6
10.4
Q)
10.2
Cl
o 10.0
_J
9.8
9.6
9.4
9.2
9.0
8.8
8.6
8.4 '---'---'---'----'----'----'-----'----''---'----'----'----'
o 2 4 6 8 10 12 14 16 18 20
Time
oX +Y

Figure 12.11 Simulation of error correction model (slow adjustment)

When [3 1 = O, solving equation 12.10 for equilibrium gives:


y* = vx* (12.12)
In this particular case:
ln( C*) = -0.11 + 0.97ln(Y*) (12.13)
Suppose that the 0.97 had indeed been 1 (as it may well be for the long-
run estímate of the consumption function). Then equation (12.13) could
be rearranged as:
C*
Y*
= e-0·11 = O.88 (12.14)

As stated above, a unitary elasticity yields a constant APC, and the value
of that APC is 0.88. But, in fact, we did not get a unitary elasticity, but
a value slightly less than one - suggesting that the APC will fall as income
rises (as the percentage increase in consumption is a bit less than that in
income ). Specifically:
C* = e--0.11 --0.03In(Y) (12.15)
Y*
So the APC depends on the level of Y. In fact, the variation is slight. As
shown in Table 12.2, at the mean level of ln(Y) the APC is 0.764, and it
Cointegration and the error correction model 411
Table 12.2 Average and marginal propensities for consumption function
y ln(Y) APC MPC e· dC MPC
(long-run) (impact)
Minimum 124 4.82 0.775 0.752 96.1 96.8 0.682
Average 198 5.29 0.764 0.741 151.5 152.2 0.672
Maximum 269 5.59 0.757 0.735 203.4 204.1 0.666
Note: Calculated from data and estimates for consumption function; details gíven in text.

ranges from 0.775 for the lowest value of Y to 0.757 for the highest. The
constant APC is therefore nota bad approximation, but to calculate its level
we do need to take account of the level of Y (i.e. the APC is not 0.88 as it
would appear to be if we ignored the intercept in the calculation).
The marginal propensity to consume will also depend on the value of
Y and is best calculated by rearranging the formula that:
elasticity = MPC (12.16)
APC
since we know the elasticity to be 0.97. The results of this calculation are
shown in Table 12.2. The MPC also varies over a small range: from 0.735
(for the lowest incomes) to 0.752 (for the highest). This MPC is the long-
run propensity - that is, an increase in income of 100 units will increase
equilibrium consumption by approximately 75 units.
The impact effect of a change in income has to be calculated from the
coefficient on the impact term. The percentage change in consumption
can be calculated as:
et _ = ebiln(Y/Y,_,)
__ (12.17)
c1-1

Table 12.2 shows the application of this formula. For each of the minimum,
mean and maximum values of Y the corresponding equilibrium C was
calculated ( = APC x Y). The percentage change given by equation (12.17),
when income is increased by one unit, is calculated, and this percentage
used to calculate the absolute increment in consumption. This increment
is the impact MPC. As expected, it is a bit, but not that much, less than
the long-run MPC. At the average level of income our results show that
an increase in income of one unit will increase consumption by 0.67 units
in the year of the rise in income. In the long run, consumption will rise
by 0.74 units, with most of the additional 0.07 units coming in the year
after the increase in income.

Exercise 12.3
Use a spreadsheet to construct a simulation of the ECM. Experiment with
different parameter values.
412 Econometrics far developing countries
Exercise 12.4
Using the data of your choice, find a non-spurious regression between two
I(l) variables and estima te and interpret an ECM.

12.5 SUMMARY OF MAIN POINTS


Regression of one non-stationary series on another is likely to yield
spurious results (i.e. inconsistent estimates). However, cointegration
analysis allows us to conduct econometric analysis of non-stationary vari-
ables. To do so the following steps should be followed:
1 Test the order of integration of each variable using the decision tree
from Chapter 10. Such analysis is often done using log values: this is
not always necessary but will usually help stabilise the variance in the
series and in many cases is more amenable to economic interpreta-
tion.
2 If the series are I(l), conduct the OLS regression as the cointegrating
regression.
3 Test the order of integration of the residuals from the cointegrating
regression using the Dickey-Fuller (DF) test, or the augmented
Dickey-Fuller (ADF) test if the DW from the DF indicates autocor-
relation. If the residuals are stationary then the dependent variable
and the regressor are cointegrated. The coefficient estimates are super-
consistent, but the standard errors invalid.
4 Define the error correction term as the residuals from the cointegrating
regression.
5 Estimate the error correction model, by regressing the differenced
dependent variable on the differenced regressor and the lagged error
correction term. (You should first check that it is valid to exclude the
in tercept.)
6 The above steps apply to the case of bivariate analysis. They may also
be applied in the case of multivariate analysis, although the procedure
then assumes that there is only one cointegrating vector, which may
not be the case. However, the test for the number of such vectors is
beyond the scope of this book.

NOTES
1 This statement is a specific form of cointegration which, more generally
defined, encompasses higher orders of integration.
2 As written here we are ignoring the intercept. The addition of a constant to
a stationary series will not alter the fact that the series is stationary, so the
omission makes no difference to the argument.
3 The critica! value depends on the number of regressors in the levels regres-
sion, but we are restricting our attention to the bivariate case.
PartV
Simultaneous equation models
This page intentionally left blank
13 Misspecification bias from single
equation estimation

13.1 INTRODUCTION

Thus far we have considered estimation of single equations. Yet in


economics many relationships are a part of a larger system of equations
(or model). If we are estimating a supply curve either price or quantity
(frequently the former) must be taken as exogenous. Usually, however,
economists would consider both variables to be endogenous (see Box 13.1
for a discussion of terminology), the system being completed with a
demand curve (and the equilibrium condition that quantity supplied equals
quantity demanded). Similarly if we analyse only the goods market in a
simple Keynesian model using the IS curve we may consider interest rates
as exogenous and the level of income as endogenous. But theory suggests
that the interest rate is also endogenous - with a further equation being
provided by the money market (the LM curve), in which the interest rate
moves to bring money demand - which also depends on income - into
line with money supply.
However, suppose that we are only interested in one of the equations
in a system. For example, the estimation of agricultura! supply schedules
is an important issue in the design of macroeconomic policy in developing
countries: structural adjustment policies rely on a reasonable degree of
price elasticity of supply from this sector. Can we then just ignore the
demand side and estímate the supply equation? As we shall demonstrate
in sections 13.2 and 13.3, the answer is no. Single-equation estimation of
a relationship that is in fact part of a larger system can lead to simul-
taneity bias. This bias arises since the endogeneity of the regressor(s) in
the system as a whole means that these regressors are related to the error
term, thus violating one of the assumptions underlying OLS.
Section 13.2 describes a method of assessing the seriousness of this bias,
and section 13.3 discusses the underlying theory. Section 13.4 is a slight
digression on the Granger and Sims tests for causality and concepts of
exogeneity. Section 13.5 moves to a discussion of the identification
problem, which must be analysed befare a simultaneous system can be
analysed. Section 13.6 concludes.
416 Econometrics far developing countries

Box 13.1 The terminology of simultaneous systems


Previous chapters have considered single equations, in which the
variable on the left-hand side is endogenous (i.e. determined within
the system) and those on the right-hand side are exogenous (i.e.
given from, or determined, outside of the model). By convention,
endogenous variables are denoted as Ys and exogenous ones as
Xs. Modern econometrics has adopted a different meaning of
exogeneity, which is discussed in section 13.5.
Once we are considering a system containing more than one equa-
tion it is quite possible that endogenous variables may appear on the
right-hand side of sorne equations. For example, consider the system:
Y1,; = !31 + f32Y2,i + f33X¡ + E¡,;

Y2,; = 'Y1 + -Y2Yu + "/3X; + Ez,;

then both Y1 and Y 2 are endogenous, even though each appears on


the right-hand side of an equation. In general, there must be the
same number of equations as there are endogenous variables. And
it is good practice to begin your analysis of a model by listing ali
the variables it contains, and noting which are endogenous and which
exogenous.
A time series model may also include the lag of an endogenous
variable on the right-hand side. Such terms are said to be pre-
determined; this expression is also applied to exogenous variables.
Even though they represent an endogenous variable the value of Y 1_
1 is given at time t, i.e. it is not to be determined within the system.
As we define them here, the exogenous variables are a sub-set of
predetermined variables. More formally, we could say that a pre-
determined variable is independent of the present and future errors
in a given equation, whereas an exogenous variable is independent
of the present, future and past errors in the equation.
The equations given above are known as the structural equations
and their parameters ( the f3s and -ys) are the structural parameters. If
we were to salve the system for Y1 and Y2, stating each only in terms
of predetermined variables, these are the reduced-form equations.
A special class of model is recursive models. A recursive model
is a multi-equation model without simultaneity: if, say, Y1 is a func-
tion of Y 2 there will be no equation in which Y 2 is a function of Y1•
For a model to be recursive, at least one of the structural equations
must be a reduced form and a model solution must be possible by
repeated substitution of one equation into another. In a recursive
model the problem of simultaneity bias does not arise as OLS yields
consistent estimates (see Chapter 14).
Simultaneity bias 417
13.2 SIMULTANEITY BIAS IN A
SUPPLY AND DEMAND MODEL
To formalise the discussion in section 13.1, suppose we wish to estimate
the supply function:
(13.1)
where Qs is quantity supplied and P is the price. But the full system has
also the demand equation and the equilibrium condition:
(13.2)

(13.3)
where Qd is quantity demanded, Q is the market-clearing equilibrium
quantity and Y is consumers' income. There are four equations (equation
13.3 is really two equations, Q = Qs and Qs = Qd) and four endogenous
variables (Q, Qs, Qd and P). Let us see what difference it makes to esti-
mate equation (13.1) first as a single equation and then taking into account
the whole model. The data used below are contained in data file SIMSIM. 1
Transforming all variables to logs and estimating by OLS yields:
A

Q~ = -8.87 + O.llP1 R2 = 0.15


(37.93) (2.62) (13.4)
where figures in brackets are t-statistics; the standard error on the esti-
mate b2 (which we will need later) is 0.042.
However, as will beco me clear in the following pages, the fact that price
is not an exogenous variable in the system as a whole means that OLS
estimation of the supply equation will be biased. In the next section we
shall derive this bias formally. Here we illustrate how to test if there is
significant simultaneity bias. To do this we have to apply estimation tech-
niques for simultaneous systems. The reader is asked to bear with us in
the application of these techniques prior to a full explanation of them,
which is given in the next chapter. For the moment, read this section to
get the basic point but it would probably pay to come back to it after
finishing Chapter 14.
The basic problem for OLS estimation is that since price is not exoge-
nous it will be related to the error term (section 13.3). Instrumental vari-
able (IV) estimation is based on choosing a variable which is exogenous
and therefore unrelated to the error term, but which at the same time may
serve as a proxy for price. (It must also be an exogenous variable that <loes
not appear in the equation whose parameters we are estimating.) In this
case we only have one choice of instrument - that is, income which is the
only exogenous variable in our system. The IV estimator of b2 is:
Iy.q.
b2!V -- - - ' - ' - (13.5)
Iyp¡
418 Econometrics for developing countries
You will find that most econometric packages allow you to select the
estimation method - if you choose IV you must specify the instruments
to be used. 2 Doing this should yield the result:
Q~ = 8.36 + 0.20P1
(30.57) (4.13) (13.6)
The price elasticity of supply now appears to be 0.2: compared to 0.1 using
OLS, this is quite a difference. Why has this difference occurred?
Here we state a result that we shall prove in the next section. We know
that if price is exogenous (and the other OLS assumptions are met) then
OLS estimation of equation (13.1) is BLUE. IV estimation will be consis-
tent, but it is not efficient (the standard errors will be larger than if using
OLS). But if price is not exogenous - as theory suggests it is not here
- then OLS will be biased and inconsistent. By contrast, IV will be con-
sistent even though the price variable is endogenous. Whether single
equation estimation is consistent or not therefore rests on the question
of whether or not price is exogenous.
The conventional approach has been to determine which variables are
exogenous and which endogenous a priori. In the supply and demand
example, we expect both price and quantity to be endogenous. Yet such
an approach conflicts with our data-oriented methodology and recent liter-
ature has been concerned to develop tests for exogeneity. But many of
these (e.g. Granger causality) do not serve our purpose here (as will be
discussed in section 13.4). The test proposed by Hausman (Hausman's
specification error test) is suitable, as it directly tackles the question of
whether or not simultaneity bias is present in the data.
The estimates obtained by OLS and IV estimation are different. The
question (as it usually boils down to in statistics) is: 'is this difference
significant?'. In other words, define the difference between the two esti-
mators as u:
U = bf' - bfLS (13.7)

We wish to test whether or not u is significantly different from zero: the


null hypothesis that u = O is equivalent to the null hypothesis that the
regressor whose coefficient we are testing is exogenous. (Why?) In this
case, u = 0.203 - 0.110 = 0.093. (Three decimal places are used in the
calculation in order to report accurately to two decimal places.)
Recall that for a t-test we deduct the hypothesised value of 13 (in this
case zero) from the observed and divide by the standard error of ¡3. The
test statistic for Hausman's is analogously defined, except that the expres-
sion is squared:
u2
m=--- (13.8)
Var(u)
which is distributed as a chi-squared with one degree of freedom.
Simultaneity bias 419
What does it mean if u turns out to be significantly greater than zero?
It meaos that the difference between IV and OLS estimation is signifi-
cant. Yet if price were exogenous the two estimation techniques should
yield the same result; the fact that they do not is symptomatic of the pres-
ence of simultaneity bias. Where the difference is significant we have
shown that the data reject the hypothesis that price is exogenous, so a
simultaneous estimation technique should be used. 3
To test whether the value of u = 0.093 obtained in the above example
is significantly different from zero we need to calculate Var(u), which
Hausman showed was given by:
Var(u) = Var(b{v) - Var(bfL 5) (13.9)
where the variances are calculated under the null hypothesis ( that price
is exogenous ). By substituting in the expressions for these variances it can
be shown that an alternative formulation for the variance is:
1 - r2
Var (u)=----;:¡- Var (bfLS) (13.10)

where r is the correlation coefficient between the variable whose


exogeneity we are testing (price) and the instrument used for the IV esti-
mator (income). For the data here this correlation is 0.908. Var(b 2°Ls) is
the square of SE(b 2°LS) ( = 0.042 2 = 0.0018). Substituting equation (13.10)
into (13.8) gives:
u 2 r2 0.093 2 X 0.908 2
m = (1 - r2) Var (b 0~ 5 ) = (1 - 0.9082 ) X 0.0018 = ·
22 57 (13.11)

This calculated value of 22.57 compares with a critica! value of chi-squared


at the 5 per cent level of 3.84, so we reject the null hypothesis that
u is insignificantly different from zero, i.e. that price is exogenous. OLS
estimation of the supply equation is therefore subject to significant simul-
taneity bias.

Exercise 13.1
Use the Indonesian national account data contained in data file INDONA
to estima te a consumption function for Indonesia using (a) O LS and (b)
IV, with investment and the trade balance as a single instrument (i.e.
define a new composite exogenous variable, equal to 1 + X - M). Use
Hausman's test to see if there is a significant difference between the two
estimators. How do you interpret this result?

An omitted-variable version of the Hausman test


Although the above test can be generalised, it becomes cumbersome when
there is more than one regressor whose exogeneity needs to be tested, or
420 Econometrics far developing countries
when we have a choice of instruments. In such cases an alternative form
of the test may be used. 4 The test involves two stages:
1 Regress each variable whose exogeneity is to be checked on all exoge-
nous variables in the model (not including any whose exogeneity is
being tested).
2 Estimate the expanded regressions, which are the simple regression
for each structural equation in turn but with all the fitted values from
step 1 included on the right-hand side. The F-test of the joint signifi-
cance of the coefficients on the fitted values in the expanded regression
is equivalent to the Hausman test.
As an example, consider the model of Maltese exports proposed by
Gatt (1995; the data are in the data file MALTA):
xs = 131 + l32P + ¡33Pw + ¡34Y + El (13.12)
xd = -y 1 + -y 2P + -y 3E + -y4CPI + -Ysl + E2 (13.13)
xs = Xd =X (13.14)
where xs and Xd are the quantity supplied and demanded of exports and
X the observed, market clearing export level; P is the export price index,
pw a world price index, Y an index of real world income, E the nominal
exchange rate, CPI the consumer price index and I is investment. All vari-
ables are logged for estimation purposes.
There are four equations here, so we should expect there to be four
endogenous variables. Desired supply and demand of exports (Xs and Xd)
and the market-clearing level of exports (X) are taken to be endogenous
- which is why we write equations (13.12) and (13.13) as we have. But
one more endogenous variable is required. The selection of this variable
is an example of what is known as model closure. We close the model by
assuming that price adjusts to equilibrate supply and demand, 5 that is,
price is taken as the fourth endogenous variable. It is this assumption we
will now test by applying the omitted variable version of the Hausman
test to see whether price is exogenous in the demand equation - the null
is that price is exogenous though our a priori expectation is that we will
reject the null.
Table 13.1 shows the regression results. The first column shows those
obtained by OLS and the second those from two-stage least squares
(TSLS, which must be used rather than IV when there is a choice of
instruments. TSLS is described more fully in Chapter 14. The technique
involves estimating the reduced form for each endogenous variable and
then using the fitted values from these regressions in the structural equa-
tions.) The variable we believe to be endogenous is P - and the coefficient
of P is very different from the two estimation procedures (as are those
on the other variables). Under OLS estimation the coefficient on price in
the demand equation is -2.88. But TSLS indicates a much stronger price
Simultaneity bias 421
effect, with a coefficient of -5.07. To test if the difference between these
two coefficients is significant we apply the omitted variable version of the
Hausman test. First P is regressed on all exogenous variables (Pw, Y, E,
CPI and !) (the results of this regression are not shown here) and calcu-
late the fitted values. We next estímate the expanded regression, that is
the demand equation with fitted P added as a regressor: 6
A

Xd = ')' 1 + ')'2 P + ')' 3 E+ ')' 4 CPI + 'Ysf + oP + E' 3 (13.15)


The results from this regression are shown in the third column of Table
13.1. The F-test on the restriction that the fitted values may be excluded
from the regression (8 = O). Since there is only one variable under consid-
eration here the t-test may be used (see Chapter 6). The null hypothesis
is that the variable is exogenous so there is no simultaneity bias. The t-
statistic from the example here is -6.37, so the null hypothesis must be
rejected. Hence price is endogenous and a simultaneous estimation tech-
nique should be used. The techniques to do so are discussed in Chapter
14; the next section examines why simultaneity bias occurs.
Table 13.1 shows the OLS and TSLS estimates of the supply equation.
The coefficients again vary between the two regressions, but it is left as
an exercise to apply the Hausman test in this case.

Exercise 13.2
Use the data in the data file MALTA to replicate Table 13.1. Test whether
price is exogenous in the supply equation.

Table 13.1 Estimation of supply and demand model for Maltese exports
Demand Supply
OLS TSLS Hausman (OLS) OLS TSLS
Constant -2.41 -2.93 -2.93 3.26 3.50
(-1.51) (-1.49) (-3.01) (3.48) (3.65)
p -2.88 -5.07 -0.02 1.91 2.21
(-4.44) (-4.81) (-0.03) (5.80) (6.12)
pw 3.81 5.62 5.62
(6.87) (6.32) (12.76)
y 0.63 1.01 1.01
(1.83) (2.29) (4.62)
E 2.70 3.05
(5.20) (5.53)
CPI -1.12 -1.38
(-2.90) (-3.35)
I 0.38 0.32
(2.65) (2.16)
Fitted P -5.05
(-6.37)
R2 0.96 0.94 0.99 0.97 0.97
Note: - indicates excluded from regression. See text for explanation of symbols.
422 Econometrics far developing countries
Exercise 13.3
Suppose that a fourth equation is added to the model of Maltese export:
(13.16)
Test the exogeneity of P and CPI in the supply and demand equations in
this expanded model.

13.3 SIMULTANEITY BIAS: THE THEORY


The supply and demand system given in equations (13.1)-(13.3) may be
solved to give the reduced form expression for price:
p = /11 - /31 + /13 y + E2 - el
(13.17)
t /32 - /12 /32 - /12 t /32 - /12
from which it may be seen immediately that price is related to the error
term in the supply equation, E 1• More specifically, multiplying equation
(13.17) through by E1 and taking expectations gives:
a2
E[(P - E(P)) ei] = - (13.18)
/32 - /12
where cr 2 is the variance of the error term. Since this latter is positive and
we would expect -y 2 to be negative (why?) then the covariance given by
equation (13.18) will be negative.
In this example we are regressing quantity (Q) on price (P). Hence as
part of the proof of unbiasedness of the OLS estimator for the supply
equation we derive the expression (lower case denoting mean deviations):
L pe 1 )
E(b2) = /32 + E ( L P2 (13.19)

If the regressor (price) is either fixed or independent of the error then


the second term on the right-hand side equals zero and the OLS esti-
mator is unbiased. But we have already shown in equation (13.17) that
price is neither of these things: it is a random variable that directly depends
on the error term. The second term will therefore not disappear and OLS
estimation is biased.
To evaluate this bias we must further manipulate equation (13.19). It
is not possible to take expectations of the second term. The expectation
of the ratio is not the ratio of the expectations. We must resort instead
to probability limits (plims; see Box 13.2 for an introduction to this
concept), for which (by Slutsky's theorem) we may separate out numer-
ator and denominator:
plim~ L pe 1
plim (b 2 ) = /32 + (13.20)
1
plim N L p 2
Simultaneity bias 423
so that the numerator will converge on the covariance between price and
the error and the denominator on the variance of price (u~). That is (using
equation (13.18)):

plim (b 2 ) = {3 2 -
1
~ (13.21)
/32 - 'Y2 <Tp

So that OLS estimation will be biased downwards - it will underestimate


the true price elasticity. We saw in section 13.2 that the OLS estimate of
the elasticity of supply was indeed lower, that is biased downward,
compared to that obtained from the consistent IV estimator. In the case
of the demand equation for Maltese exports there is an upward bias; it
is left as an exercise to examine why this is so.

Box 13.2 Probability limits (plims)


An estimator is said to be consistent if its distribution converges on
the population value as the sample size increases. We write that b
is a consistent estimator of 13 as:
plim(b) = 13
Sufficient conditions for an estimator to be consistent are that (a)
it is asymptotically unbiased, that is, its expected value tends to the
population value as the sample size tends to infinity; and (b) that
the variance of the estimator tends to zero as the sample size
increase. Property (b) ensures that the distribution of the estimator
converges as the sample size increases and property (a) means that
convergence is on the population value.
Hence, if an estimator is consistent it may be used to estimate
population values for large samples. But in small samples a consis-
tent estimator may well be biased and so should not be used.
The algebra of plims has an advantage over expectations since the
plim of the products is equal to the product of the plims, which is
not true for expectations. That is:
plim(xy) = plim(x) + plim(y)
. (....:!__) _ plim(x)
pl im Y - plim(y)

This property of plims is known as Slutsky's theorem.

We have just stated again that the IV estimator is consistent, and we


are now in a position to illustrate this. From equation (13.5) it follows
that:
424 Econometrics far developing countries

E(b~) = /32 + .L
y¡ ei ¡ (13.22)
LY¡P¡

We cannot evaluate the final expectation for the same reason as befare.
But if probability limits are taken, then, since income is independent of
the error term, we know that plim(l/N)¡yE 1 = O. Hence plim(b21v) = [3 2,
showing that instrumental variables gives a consistent estimator. Note that
the estimator is consistent, but not unbiased, as the final expression in
equation (13.22) cannot be reduced to zero. Instrumental variable estima-
tion and TSLS are for this reason 'large sample techniques', the distribu-
tion of the estimate converges on the population value as the sample size
increases - so estimation from a small sample may plausibly give an esti-
mate quite far removed from this population value. 'Large sample' should
ideally be taken as at least 50 observations, though we often do not have
so many observations for developing countries (notably with time-series
data). But the point is an important one, and it would be pointless to apply
these techniques to samples of fewer than 25 observations.
The bias in OLS estimation of the supply equation arase because price
is not exogenous as OLS assumes. The Hausman test is based on testing
whether or not OLS (which will be inconsistent if price is exogenous)
gives a significantly different estimate to a technique which is consistent
when there is a problem of simultaneity (we used IV, but, as we shall
see in Chapter 14, there are other techniques). If we find that, contrary
to our theoretical expectation, there is no significant simultaneity bias,
then the OLS estimates may be used. Whilst the IV results are the same as
(insignificantly different from) those obtained by OLS (both methods
are consistent when there is no simultaneity bias), the former are less
efficient.
Hausman's test is a test of exogeneity that is directly related to the
problem in hand: is OLS estimation appropriate or not? In the next section
we discuss other definitions and tests for exogeneity that are common in
the literature but which do not have the same intuitive appeal.

Exercise 13.4
The consumption function is often estimated as:
et= f31 + f32Y1 + E¡ (13.23)
Yet income also depends on consumption through the accounting identity:
(13.24
where I is investment and (X - M) the balance of trade in goods and
services, both of which are taken to be exogenous. 7 Show that single-equa-
tion estimation of the consumption function gives a biased estimator of
the MPC, with an upward bias of:
Simultaneity bias 425

plim (b 2) - /3 2 = -
1
-
1 - /32
~
<Iy
(13.25)

Did your results from exercise 13.1 conform with this expression?

13.4 THE GRANGER AND SIMS TESTS FOR CAUSALITY


AND CONCEPTS OF EXOGENEITY
The approach of orthodox econometrics to simultaneous modelling has
been to specify the exogenous and endogenous variables a priori: the view
is that exogeneity cannot be tested and must be assumed. In fact, the
Hausman test presented in section 13.2 can tell us if assuming a variable
to be exogenous or endogenous makes any significant difference to our
regression results - which is what we need to know - and so is a suitable
test for determining exogeneity from the data.
Modern econometrics has evolved other definitions of exogeneity, one
of which involves the concept of Granger causality. This technique, and
the related definitions of exogeneity, are presented here, as the reader is
very likely to come across them in applied work. But it is important to
realise that Granger causality is not about causality in the normally
accepted sense of the word - Leamer has been suggested that 'prece-
dence' would be a more appropriate term to describe what the Granger
test captures.

Granger causality
The Granger test that X does not Granger cause Y is the F-test that the
Xs may be excluded from the equation:
k k
Yt = 130 + I 13i Yt-i + I "{¡X¡_¡+ E¡ (13.26)
i=l i=l

Thus applying the test requires three steps:


1 Estimate the unrestricted model given by equation (13.26); the length
of the lag used depends on (a) how much data you have available;
and (b) examining the data. It is advisable to check that the result is
not sensitive to the lag chosen.
2 Estimate the restricted model by regressing Y just on the lagged Ys.
3 Test the restriction with the F-test in the usual way (see Chapter 6).
The null hypothesis is that X does not Granger cause Y.
This procedure can then be repeated by 'swapping the variables' to inves-
tigate if Y Granger causes X. There are thus four possible outcomes: (a)
no causal relationship between the two variables; (b) unidirectional
causality from X to Y; (c) unidirectional causality from Y to X; and (c)
bidirectional causality (X causes Y and Y causes X). As an example we
426 Econometrics far developing countries
apply the test to the price and quantity data from the model of Maltese
exports used in section 13.2. First, to test if quantity Granger causes price
we estimate the unrestricted and restricted models:
pt = !31 + !32Pt-l + ¡33Pt-2 + 'Y1º1-l + 'Y2ºt-2 + E1,1 (13.27)
1
pt = !31 + !32Pt-l + ¡33Pt-2 + E 1,t (13.28)
The results are shown in Table 13.2, where the RSS can be seen to have
nearly doubled as a result of imposing the restriction. The calculated F-
statistic is 8.55, compared to a critical value of 3.49 at the 5 per cent level,
so we reject the null hypothesis. The data suggest that quantity does
Granger cause price.
What does this result really tell us? If changes in quantity are to cause
changes in price, then the quantity change must come befare (precede)
the price change - if the quantity change comes after the price change
then it cannot be said to have caused it. The Granger test therefare sees
if changes in quantity have a significant impact on future prices. However,
whilst it is true that if event X fallows event Y then X cannot have caused
Y, it does not fallow that if X precedes Y then X has necessarily caused
Y in the normal sense of the term. That X comes befare Y is a neces-
sary, but not sufficient, condition far Y to be caused by X.
Table 13.2 also shows the results from testing if price Granger causes
quantity. The calculated F-statistic is 0.63, so we accept the null hypoth-
esis that price does not cause quantity. To reject the result would suggest
bidirectional causality, which may indicate that a third variable is in fact
determining changes in both price and quantity. However, our findings
indicate unidirectional causality from quantity to price.

Table 13.2 Granger and Sims test for Maltese export quantity and prices
Granger Sims
p p X X p p X X
e 0.41 0.14 0.22 0.42 1.37 1.43 -1.93 -1.65
P_z -0.65 -0.55 -0.13 0.31 1.06
P_¡ 1.31 1.52 0.27 0.72 -1.22
p -0.53 1.77
p+l 0.15
P+z 0.99
X_z -0.11 -0.10 -0.04 0.11 0.12
X_¡ 1.31 0.96 0.98 0.30 0.28
X 0.04 0.16
X+ i 0.11
X+z 0.01
n 25 25 25 25 23 23 23 23
Rz 0.98 0.97 0.98 0.98 0.93 0.93 0.98 0.93
RSS 0.083 0.154 0.253 0.269 0.301 0.305 0.270 0.782
Notes: - indicates excluded from the regression. See text for symbols and explanation.
Simultaneity bias 427
The Sims test
The test proposed by Sims is differently specified, but has the same intu-
itive interpretation as the Granger test. The unrestricted equation for the
Sims test is: k2
Yt = f3o + I f3;'i+j + E (13.29)
j=kl
that is, Y is regressed on lagged, current and future values of X (the length
of the lag and the lead need not be equal). The restricted equation
excludes the lagged values of X:
o
Y 1 = [3 0 + I [3jXt+j + E (13.30)
j=kl

Be sure to use the same sample size! So we are testing the null that the
coefficients on the lead terms are jointly zero: i.e. future values of X do
not affect Y. The null hypothesis is that Y does not cause X. Note the
difference between the Sims and Granger tests. Here Y is the dependent
variable, but we are checking if Y causes X, which is the opposite of the
Granger test. If future X is significant, then Y cannot cause X as it does
not precede it.
Table 13.2 reports the results for the Sims test using the price and quan-
tity data. First, price is regressed on past, present and future prices. The
omission of the lead terms barely changes the RSS, so that the calculated
F-statistic is only 0.11. Hence we accept the null hypothesis that price
does not cause quantity. By contrast, the calculated F-statistic for the
hypothesis that quantity does not cause price is 16.12, so that the null is
rejected. We find that quantity does cause price. These findings are, of
course, the same as those obtained with the Granger test.

Exercise 13.5
Use the data on money and prices in Tanzania (date file TANMON) to
test for Granger causality between the two variables. Carry out also the
Sims test and compare your results.

Exogeneity
Three concepts of exogeneity may be identified: weak, strong and super. 8
A variable, X, is weakly exogenous if there is no loss of information by
analysing the distribution of Y conditional upon X (which is what OLS
regression does) and ignoring the stochastic behaviour of X itself. If our
concern is to establish the appropriate estimation technique, then weak
exogeneity is all which need concern us. We have already seen that the
Hausman test can provide us with this information. Granger causality, on
the other hand, is neither necessary nor sufficient to establish weak
exogeneity.
428 Econometrics far developing countries
Granger causality enters the picture because the definition for X to be
strongly exogenous in a model containing X and Y is that X should be
weakly exogenous and the X should not be Granger caused by Y. Super-
exogeneity is related to the Lucas critique (Lucas, 1976), which is the
notion that behaviourial relationships can change in the face of policy
changes (i.e. adjustment in 'exogenous' policy variables). A variable, X,
is said to be super-exogenous if (a) X is weakly exogenous; and (b) model
parameters are invariant to changes in X (formally speaking, the marginal
distribution of X). See, for example, Maddala (1992) for a more extended
discussion of these concepts. Our main point is to emphasise that the
Hausman test comes closest to the conception of exogeneity required in
the context of estimating simultaneous equations.

13.5 THE IDENTIFICATION PROBLEM


Sections 13.2 and 13.3 showed that estimation of a single equation by OLS
may be inappropriate if the equation is a part of a system of equations.
The obvious next step might seem to be to discuss what estimation tech-
niques should be applied in these circumstances. However, before we can
do that we must first deal with the identification problem. This is not the
digression it might at first appear. The appropriate technique for esti-
mating an equation in a simultaneous system depends on the identification
status of that equation: if an equation is not identified then its parame-
ters cannot be estimated.

ldentification in the supply and demand model


The supply and demand system presented in section 13.2 (equations
(13.1)-(13.3)) may also be presented graphically, as is done in Figure 13.1,
where a deterministic form of the model is presented (both error terms
are zero). Changes in the endogenous variables are given by movements
along the schedules, whereas a change in an exogenous variable shifts the
schedules in which it appears. In this case there is only one exogenous
variable - income, and the demand schedule will (for a normal good) shift
to the right as income increases. Figure 13.1 shows the demand schedule
for three different periods in which there are three different levels of
income. What can you say about the three resulting equilibria?
The supply schedule is fixed, whereas the demand curve is shifting as
income changes. All observed equilibria therefore lie upon the single
supply schedule - the schedule is sketched out by the different observa-
tions. The same is not true for the demand curve. The observations yield
no information as to the slope of the demand curve: there is no estima-
tion technique to estímate the demand curve under these circumstances.
The demand curve is said to be under- or not identified, whereas
the supply curve is identified. An equation is identified if it is possible to
Simultaneity bias 429
p s

o
Figure 13.1 The identification of the supply curve by the demand curve

estimate its parameters. lf an equation is unidentified the parameters


can only be estimated by respecifying the model. (Though we should
be wary of model respecifications which are made only to achieve iden-
tification!) Even though we cannot estimate the model parameters, if
we know the direction of the simultaneity bias then OLS estimation will
give us an upper or lower limit - an application of this fact is given in
exercise 13.9.
Suppose instead that the supply schedule also contains an exogenous
variable, W, which is a proxy for weather conditions:
Q~ = f31 + f32Pt + f33Wt + El,t (13.31)
= O, the demand curve will be fixed and the shifts in the supply curve
If -y 3
will sketch out the demand curve. Now the demand curve will be identi-
fied (we can estimate its parameters) and the supply curve will not - this
is the case shown in Figure 13.2. Of course, if neither exogenous variable
enters the model (f3 3 = -y 3 = O) then neither curve is identified: we cannot
estimate any of the parameters in the model. But in the more general
case when there is an exogenous variable in each schedule which is
excluded from the other schedule, then both equations are identified.
This example suggests what is required for an equation to be identi-
fied: it must exclude exogenous variables that appear elsewhere in the
model. Though we cannot present this point in the visually appealing
manner of the two-equation model above for larger systems, the intuition
430 Econometrics far developing countries

Figure 13.2 The identification of the demand curve by the supply curve

generalises and is formally embodied in the rank and order conditions for
identification. These conditions are the subject of the next section.

The rank and order conditions


The rank and order conditions may at first seem complicated. But with
sorne practice their application becomes straightforward. To explain them
we must first define:
M: the number of endogenous variables in the system
K: the number of predetermined variables in the system (including the
intercept)
m: the number of endogenous variables in the equation under consider-
ation
k: the number of predetermined variables in the equation under consid-
eration
In applying the rank and order conditions the first step is to work out M
and K and m and k for each equation we wish to estimate. (As a general
rule, whenever working with a simultaneous model under any circum-
stances this is a useful starting point.)
For the supply and demand model in equations (13.1)-(13.3) we have
four endogenous variables, QS, Qd, Q and P, that is, M = 4. There are
Simultaneity bias 431
Table 13.3 Coefficient matrix for supply and demand model
Equation Constant Q Qs Qd p y
Supply f31 o -1 o f32 o
Demand "11 o o -1 "12 "{3
Equil. 1 o 1 -1 o o o
Equil. 2 o o -1 1 o o

two exogenous variables, Y and the intercept; i.e. K = 2. We wish to estí-


mate the supply and demand equations:
Supply equation: m = 2 and k = 1
Demand equation: m = 2 and k = 2
The two equations Q = Qs and Qs = Qd are identities. We do not have
to test identities to see if they are identified. This is immediately apparent
if we think about why we are looking at the identification problem in the
first place. We are doing so to determine whether or nor we can estímate
the parameters in an equation - yet in an identity there are no parame-
ters to estimate, so there is no problem.
To apply the rank condition we must write out the coefficient matrix
for all the equations in the system (including identities), with all variables
(except the error) on the right-hand side. This is usually done in tabular
form: Table 13.3 shows the coefficients for the supply and demand model
of equations (13.1)-(13.3) (i.e. excluding weather in the supply equation).
The rank condition is as follows: Por an equation to be identified it
must be possible to construct at least one (M - 1) x (M - 1) matrix with
a non-zero determinant from the coefficients of those variables excluded
from that equation but included in other equations in the model.
To apply this condition in practice, cross out the line for the equa-
tion under consideration and all the columns which do not have a zero
coefficient in the equation under consideration; then write down the coef-
ficients that are left - these are the coefficients 'of those variables excluded
from that equation but included in other equations in the model'.
Por the supply equation we get the matrix:

(o~ -~ ~º)o
1
(13.32)

from which we must form a non-zero determinant of 3 x 3. In this case


there is only one possible determinant. In general, however, though the
resulting coefficient matrix must have (M - 1) rows (why?) it may have
more (or less) than M - 1 columns (why?) and so there will be a choice
of matrices (or none at all) that can be formed. If there is a choice, only
one of the possible combinations need have a non-zero determinant for
the rank condition to be satisfied.
432 Econometrics for developing countries
The determinant of the matrix formed by the coefficients of vari-
ables excluded from the supply equation is -y 3 , so that the supply curve
meets the rank condition for identification so long as -y 3 is not zero. This
makes intuitive sense, if -y 3 = O, then income <loes not appear in the
demand curve and so it will not shift to sketch out the supply curve -
which, as we saw in the previous section, means that the equation will
not be identified.
When we come to apply the rank condition to the demand equation
we find there are only two columns with zeros so the resulting matrix is
2 x 2 - it is not possible to find any determinant from a 3 x 3 matrix (let
alone see whether it is zero ). Therefore the demand equation <loes not
satisfy the rank condition and is not identified - it is not possible to estí-
mate its parameters by any technique. We re-emphasise that we do not
analyse identification for identities.
The rank condition is necessary and sufficient for identification, but it
<loes not yield all the information we require - it <loes not help to choose
the appropriate estimation technique. We must also apply the arder condi-
tion, which may be summarised as follows:
if K - k < m - 1 the equation is under-identified
if K - k = m - 1 the equation is just identified
if K - k > m - 1 the equation is over-identified
Far the supply equation K - k = 1 and m - 1 = 1, so that the equation
is just identified. Even though the rank condition is a necessary and suffi-
cient condition for identification (whereas the arder condition is only
necessary) we use the arder condition to find out if an equation in just
or over-identified since this affects how the equation is estimated. 9
As we said at the beginning of this section, this is not all as compli-
cated as it seems at first, as is apparent from doing a few examples. So
let us turn to a second example. Consider the Keynesian macro model
described in the following five equations:

Y =C+l+X-M (13.33)
C = 131 + 132 y + El (13.34)

M = 'Y1 + "12 Y + E2 (13.35)

Md = <!>1 + <l>2Y + <j>3r + E4 (13.36)

] = 81 + 82r + E3 (13.37)
Md = Ms = L_
p (13.38)

where Y is income, C consumption, 1 investment, X exparts, M imports,


r the real interest rate, Ms ( = LIP) real money supply, Md money demand
Simultaneity bias 433
Table 13.4 Coefficient matrix and rank and order conditions for Keynesian
macro model
Equation Const. y e I X M r VP Rank m-1 K - k Identified
condition
satisfied
ldentity o 1 1 o
-1 -1 -1 o n.a. ldentity
Consumption 131 132 -1 o oo o o Yes 1 2 O ver
Imports 'Y1 'Y2 o o o -1 o o Yes 1 2 O ver
lnvestment º1 o o -1 o o º2 o Yes o 1 O ver
Money market <1>1 <1>2 o o o o <1>3 -1 Yes o o Just
Note: Model values: M = 5, K = 3

and the E¡S are error terms. We combine equations (13.36) and (13.38)
so that there are five equations and five endogenous variables (Y, C, I, M
and r) and three exogenous variables (L/P, X and the intercept). Table 13.4
gives the coefficient matrix. The final columns also give the information
necessary for the arder condition, with the last column summarising the
identification status of each equation. A table such as this provides sys-
tematic means of working through the identification process.
We do not need to discuss the identification of the national accounting
identity - we already know all the coefficients are exactly one (or minus
one, as the case may be). Let us consider the consumption function. The
matrix of coefficients of variables excluded from this equation is the 4 x
5 matrix:

(-~ -~ -~
-1 o o
o o o
~ ~)
82
~3
o
-1 (13.39)

from which we must form one 4 x 4 matrix with a non-zero determinant.


Let us try the first the 4 x 4 matrix with no unknown parameters, i.e. no
¡3s, -ys, etc., since this will be unambiguously zero or not - that is, its value
will not depend on the value of any of the model's parameters):

(~~ -~ -i ~)
o o o -1 (13.40)

The determinant of this matrix may be calculated as -1, so that the


consumption function satisfies the rank condition - we know it to be iden-
tified. But we must apply the arder condition to see if it is just or
over-identified. It tums out to be over-identified (m - 1 = 1, K - k = 2).
434 Econometrics for developing countries
The import and investment functions are similarly over-identified, whereas
the money market equilibrium condition is just identified.

Exercise 13.6
Replacing equation (13.1) by equation (13.31) (i.e. use the supply equa-
tion with weather), test for identification in the supply and demand model.

13.6 SUMMARY OF MAIN POINTS


1 Most economic models consist of more than one equation. The esti-
mation of single equations from a multi-equation model by OLS can
result in simultaneity bias, as right-hand side variables which are
endogenous to the model will be correlated with the error term.
2 The Hausman test allows us to test if there is significant simultaneity
bias present in the data. If there is, then OLS is not the appropriate
estimation technique.
3 Two versions of the Hausman test are presented. The omitted vari-
ables version should be applied when there is a choice of instruments
or more than one variable whose exogeneity is being tested.
4 Severa! concepts of exogeneity are used in the econometric literature.
One of these, strong exogeneity, is related to Granger causality, which
is a test of whether X precedes Y. However, only weak exogeneity, as
tested by the Hausman test, need be analysed if the concern is to
establish the appropriate estimation technique.
5 If the parameters of an equation may be estimated the equation is
said to be identified. If an equation is under-identified then there is
no technique by which estimates of the parameters may be obtained;
the only option is to respecify the model (though estimates of upper
or lower limits may be obtained, which may be useful). Identification
of an equation is determined by the rank and arder conditions. The
rank condition is necessary and sufficient - if it is not satisfied then
the equation is under-identified. If the rank condition is met then the
arder condition is applied to discover if the equation is just or over-
identified. This latter step is necessary to determine the appropriate
means of estimation, which are the subject of Chapter 15.

ADDITIONAL EXERCISES

Exercise 13.7
Use the Indonesian national accounts data (date file INDONA) to test
whether income is exogenous in the consumption and import functions in
the simple Keynesian model:
Ct = 131 + 132 yt + El,t
Simultaneity bias 435
M, = "Y1 + "Y2Yt + E2,t

Y=C+I+X-M
where Mis imports, C consumption, Y income, I investment and X exports.

Exercise 13.8
Using equations (13.1) and (13.2) illustrate that the OLS estimate of
the coefficient on price suffers from upward bias. (The algebra for this
question is a bit messy.) Does this finding agree with the results shown
in Table 13.1?

Exercise 13.9
In the demand and supply model in equations (13.1)-(13.3) the demand
equation is unidentified, so no simultaneous technique may be applied to
estimation of the price elasticity of demand. But of course single-equa-
tion estimation by OLS is still possible. Carry out this regression and
comment on the information given by the coefficient.

Exercise 13.10
Comment on the estimation procedure for the parameters in the following
model:

HEXP¡ = 131 + l32Y; + E2,;

where LE is lite expectancy at birth, y income per capita and HEXP


health expenditure per person.
What different interpretations should be placed on the coefficient of
income in the regression for lite expectancy as between the structural
equation and the reduced form?

NOTES
1 The data have been generated by a simulation of the model specified in equa-
tions (13.1 )-(13.3).
2 The supply equation actually has two regressors - price and the intercept term.
Sorne packages, e.g. Microfit, require you to specify as many instruments as
regressors including the intercept term. If this is the case simply include the
intercept in your list of instruments - it will act as its own instrument.
3 Subject to comments made in the next section.
4 A version of the omitted-variable Hausman test was used in Chapter 10 as an
equivalent form of the Plosser-Schwert-White differencing test.
5 An alternative closure would be quantity adjustment in which the actual leve!
of exports is constrained to the mínimum of supply and demand, i.e. X =
min¡xs,xd).
436 Econometrics far developing countries
6 lf we were testing the exogeneity of more than one variable then the fitted
value for each of these variables would be included in the expanded regres-
sion.
7 The consumption function is the 'standard text book' example of simultaneity
bias. We did not use it as the accounting identity is usually presented in
for the case of a closed economy. We have included the trade balance since
we use actual data; there are no real economies for which it is true that Y =
e+ 1.
8 The reader is warned that exogeneity is a term for which differing definitions
abound. We present here those suggested by Engle et al. (1983) which have
the widest currency.
9 It is perfectly possible to check first the order condition and then the rank.
There is no strong argument either way. An equation that fails the order condi-
tion must fail the rank condition (why?) but not vice versa. Doing the rank
condition first may therefore save a small amount of time. In Chapter 14 it
will also be seen that it is possible to apply the technique of two- (or three-)
stage least squares to both just and over-identified equations: if this is the
intention then the order condition is superfluous.
14 Estimating simultaneous
equation models

14.1 INTRODUCTION
The previous chapters have shown that OLS estimation may be inappro-
priate for an equation which is part of a system of equations and that it
may not be possible to estimate the parameters of sorne equations at all
(if the equation is unidentified). We also saw that an equation may be
just or over-identified and said that this affects how the equation may be
estimated. This chapter discusses the appropriate estimation techniques
for these different cases.
We begin, however, in section 14.2 with a presentation of recursive
systems as a special class of multi-equation model. Section 14.3 discusses
the method of indirect least squares (ILS) which may only be used for
equations which are just identified. Section 14.4 presents instrumental vari-
ables and the related method of two-stage least squares. These techniques
are called limited information estimation as they estimate the parameters
of one equation at a time. Section 14.5 employs the various techniques to
estimate a consumption function for Indonesia.
In section 14.6 we discuss seemingly unrelated regressions (SUR) and
three-stage least squares (3SLS) which is a full information technique -
all the models' (estimable) parameters are estimated simultaneously.
Section 14.7 provides a summary of the chapter's main points.

14.2 RECURSIVE MODELS


A recursive model is a multi-equation model without simultaneity, it may
be written in general form as:
Y1,; = ª1 + azX¡ + El,i
Y2,i = ~1 + ~zX¡ + ~3 Y1,i + E2,i

Y3,i = 'Y1 + 'Y2X; + 'Y3 Yl,i + 'Y4 Y2,i + E3,i

(14.1)
438 Econometrics for developing countries
where the Ys are endogenous and X, which may be a vector, exogenous.
The problem of simultaneity bias arises since one of the regressors is cor-
related with the error term. This problem will not affect X as it is exoge-
nous. But in a simultaneous model - in which say Y1 depends on Y 2 and Y 2
depends on Y1 - then a variable which is the dependent variable in one
equation will be the regressor in another and simultaneity bias will be
present. But this problem does not appear in the recursive model. The
expression far Y1 is equation (14.1) in a reduced farm, so clearly OLS can
be used. Since Y1 is a function of only X and its own error E 1, it will not be
related to any of the other error terms appearing in the model. The reduced
farm far Y 2 shows it to be a function of X and E1 and E2• So Y2 will not be
correlated to, far example, E3 when it appears as a regressor in the equa-
tion of Y 3, nor E4 and so on. The same argument carries through far each
of the endogenous variables. As there is no simultaneity bias the structural
equations from a recursive model apparently may be estimated by OLS.
In fact there is a problem here, which we shall state rather than demon-
strate. In a recursive model the arder condition is satisfied far only the
first equation. However, the other equations may be shown to be identi-
fied, provided there is no cross-equation correlation between the error
terms, i.e. E(E;Ej) =O far all i =I= j. However, OLS is consistent rather than
unbiased, and so only valid far large samples.
There are two important exceptions to the above argument. The first
should be apparent from the preceding paragraph, i.e. what happens when
the error terms are cross-correlated, e.g. E(E1E2) =I= O. In this case multi-
plying the expression far Y1 through by Ez shows that E(yE 2) is not zero,
so that OLS estimation of the equation far Y 2 is now biased and incon-
sistent. In fact, cross-correlation of the error terms leads us away from
OLS even when the equations do not apparently farm part of a multi-
equation model - that is, they are what is known as seemingly unrelated
regressions. The appropriate estimation technique under these circum-
stances is discussed in section 14.6.
We have discussed here a general farm of the recursive model. In
economics such models are most likely to arise on account of lags in
behaviourial relationships. Examples of this fact are provided in the exer-
cises. But this case is the second exception to our initial description of
using OLS far the recursive model. If the first equation contains a lagged
endogenous variable, then estimation of this equation by OLS is consis-
tent but not unbiased. This result fallows since simple substitution can
obtain a reduced farm in which Y1 is a function of its own lag, under
which circumstances we know OLS to be biased but consistent.

Exercise 14.1
In a simple two-equation Keynesian model of the closed economy, the
accounting identity:
Estimating simultaneous equation models 439
(14.2)
may be combined either with a contemporaneous consumption function:
(14.3)
or one in which the effect of income on consumption operates with a lag
(14.4)
Show that the model with the lagged consumption function is recursive,
whereas that with the contemporaneous function is not. Discuss the impli-
cations of this finding far estimation of the marginal propensity to
consume.

Exercise 14.2
Consider the five-equation IS-LM model given in equations (13.33)-
(13.37). Show that a lagged consumption function does not make the
model recursive. Show that the model is recursive if the income term is
lagged in both the import and consumption functions and that investment
depends on the lagged interest rate. In the latter case explain the arder
of substitution to obtain a recursive solution to the model.

14.3 INDIRECT LEAST SQUARES


We start, once again, with our familiar supply and demand model:
Qt = f31 + f32Pt + E1,1 (14.5)
Qf = 'Y1 + 'Y2Pt + 'Y3yt + E2,1 (14.6)

Q = Qf = Qf (14.7)
where P and Q are price and quantity, the s and d superscripts denote
supply and demand, and Y is income.
This model may be solved to get the fallowing reduced-farm expres-
sions far price and quantity:
p = V1 - /31 + ')13 y + 82 - 81
(14.8)
t /32 - 1'2 /32 - 1'2 t /32 - 1'2

Q = /321'1 - /311'2 + /321'3 y + f32e2 - 1'2B1


(14.9)
t /32 - 1'2 /32 - 1'2 t /32 - 1'2

which we may write as:


Pi = 'TT1 + '1T2Y1 + w1,1 (14.10)

Qt = 'TT3 + 'TT4Y1 + w2,t (14.11)


440 Econometrics far developing countries
where
'Y1 - /31 - ')13
7T1 = 7T2 -
/32 - 'Y2 /32 - 'Y2

/321'1 - /311'2
7T3 =
/32 - 'Y2

(14.12)

Since these are reduced-form expressions the variables on the right-hand


side are, by definition, exogenous. Equations (14.10) and (14.11) may
therefore be estimated by OLS. Then we can work back from the reduced
form coefficients (the 'lTS) to the structural coefficients (the ~s and -ys) -
a technique known as indirect least squares, since we obtain the required
parameters indirectly by estimating the reduced form equations. But we
can see from the expressions gathered under equation (14.12) that we
only have four equations (the estimates for 'lT1, 'lT2 , 'lT 3 and 7T4) for five
unknowns (-y 1, -y 2 , -y 3, -y 4 and -y 5).
This situation of insufficient equations has arisen because the demand
equation is unidentified. We have already said that it is not possible to
estimate the parameters of an unidentified equation - here we have a
concrete example of this impossibility in practice. But we have also said
that we should be able to get estimates for equations which are identi-
fied, in this case the supply equation. This is indeed the case, since a little
manipulation shows that:

(14.13)

So we can obtain estimates of the parameters of the supply equation,


which is just identified, but not for the demand equation, which is uniden-
tified.

Exercise 14.3
In the following model (where W is an index of weather conditions):
Q 5 = ~1 + ~2p + ~3 W + El (14.14)
Qd = 'Y1 + 'Y2P + Ez (14.15)
Qs = Qd = Q (14.16)
which parameters may be estimated using ILS? Derive the algebraic
expressions to calculate these parameters. Discuss your answer in the light
of the identification of the equations.
Estimating simultaneous equation models 441
To apply ILS to our supply and demand example (equations (14.5)-
(14.7)) we first estimate the reduced-form equations. We do this using the
data set in data file SIMSIM, where all series have been subject to the
log transformation:
A

P1 = -1.062 +l.188Y1 (14.17)


Q 1
= 8.141 + 0.241Y1 (14.18)
where we have retained three decimal places for purposes of calculation.
Using the formulae in equation 14.13 we have: 1
/\

biLS = ~4 = 0.241 = 0. 204 (14.19)


7rz 1.188

bfLS = ~3 - b2;1 = 8.141 -7- 0.203 X 1.062 = 8.36

If you refer back to Chapter 13, you will see that these are not the
same as the results which were obtained by OLS (by which the price elas-
ticity of supply is estimated as 0.11). The latter are biased and inconsistent,
whereas the ILS estimates are consistent; but they are not unbiased. 2 That
is, in common with all techniques for estimating equations in simultaneous
systems, ILS is a large-sample technique. Small-sample bias may be quite
serious. Developing-country data sets often do not permit very large
samples but a sample size of less than 20 would cast serious doubt on the
validity of your results and it is really preferable to work with at least 25
observations.
The above example shows how ILS may be applied to the estimation
of an equation which is just identified. What about when an equation is
over-identified? We have stated that ILS is then inappropriate, and it
is easy to show why this is so. Suppose that our demand equation has an
additional exogenous variable, A, which is the real value of advertising
expenditure per capita:
(14.20)
Checking the identification of the behaviourial relationships in this model
(which is left as an exercise) shows that, whilst the demand equation
remains unidentified, the supply schedule is now over-identified.
The consequences of this over-identification are apparent when we write
out the reduced form equations:
(14.21)

(14.22)
We will now get estimates of six 7TS, but there are only five structural
coefficients to be estimated (13 1, 132 , "Y 1, "Y 2 and "'1 3). We have too many
442 Econometrics for developing countries
equations, so there will be no unique solution. There will be more than
one solution for the parameters of the supply curve (the ¡3s) and no means
of choosing between them as to the 'right one'. ILS therefore cannot be
used when an equation is over-identified.
In practice, ILS is little used even if an equation is just identified. This
neglect is because of the algebra (which rapidly becomes tedious in large
models) and subsequent substitution required to get the structural coef-
ficients from the reduced-form estimates. The method mainly survives as
an introduction to the identification problem in many texts. The preferred
method is instrumental variable estimation. In the next section we deal
with this technique first for just identified equations and then in its gener-
alised form (for equations which are either just or over-identified), which
is equivalent to two-stage least squares.

Exercise 14.4
The Human Development Index may be modelled as a function of income
per capita and population growth. But population growth itself is a function
of the HDI. Which parameters in this model may be estimated? Derive
the algebraic expressions to estimate them by ILS. Calculate these esti-
mates using the data in the data file SOCECON.

14.4 INSTRUMENTAL VARIABLE ESTIMATION AND


TWO-STAGE LEAST SQUARES
The method of instrumental variables was briefly introduced in Chapter
13. IV gets around the simultaneity bias caused by the relationship
between the regressor and the error by using a proxy which is unrelated
to the error but is correlated to the variable whose coefficient we wish
to estimate. The choice of proxy is, in practice, circumscribed by the
variables in our data set. Recall that the exogenous variable chosen
must be one that does not appear in the equation whose parameters
we are estimating.
Consider, once again, the demand and supply example, in which supply
is a function of price and demand of price and income. We have estab-
lished that it is not possible to estimate the parameters of the demand
function as it is under-identified. We can see that here, by asking what
proxy can we use for the endogenous regressor, price, in the demand
equation. We cannot use income, since that already appears in the equa-
tion. The fact that there are no variables in the model excluded in the
demand schedule means that it cannot be estimated (the parallels with
the identification conditions should be clear). For the supply schedule, on
the other hand, we may proxy price by income. This was done already in
Chapter 13 where we obtained the result (t-statistics in parentheses in this
and subsequent equations in this section):
Estimating simultaneous equation models 443
Q/ = 8.36 + 0.20Pt
(30.57) (4.13) (14.23)
We will not discuss the derivation of the IV estimator here, as it may be
calculated by a statistics or econometrics package. 3 More important is to
understand its appropriate use and interpretation. 4
You will have noticed that the results in equation (14.23) are the same
as those obtained by ILS, reported in equation (14.19). Provided an equa-
tion is just identified (which is when they should be used) then IV and
ILS will always yield identical results - they are equivalent. This equiva-
lence is easy to demonstrate:

bILS = ~ = ~yq/~y2 = ~yq = b!V


(14.24)
2 7Tz ~pyl~y2 ~py 2

And like ILS, IV estimates are consistent but not unbiased; it is also
a large sample estimator. This must obviously be so since it has just been
demonstrated that the two estimators are algebraically identical.
The R 2 obtained by OLS will always be greater than that resulting from
IV estimation. Indeed, the latter can be negative, which is symptomatic
of poor model specification, and perhaps that the equation is not identi-
fied at all.
What if we have a choice of instruments. Which should we choose? The
answer is that this problem will not arise: if we have a choice of instru-
ments then the equation will be over-, not just, identified and the use of
IV with a single instrument is inappropriate. Rather, we should apply
2SLS which utilises the full range of instruments.
The supply and demand model, with advertising included in the demand
schedule, is an example of this situation. As before, the demand equation
is unidentified. The supply equation is over-identified, and we have a
choice of instruments: either income or advertising. Estimating the model
using advertising as the instrument gives:
A

Qt = 7.32 + 0.39
(1.19) (0.35) (14.25)
which is a rather different answer to that provided by using income (here
price is not significant), given in equation (14.23). There is no basis for
saying that either equation (14.19) or (14.21) is the 'right answer'. This
situation is analogous to that in IV estimation, where we had more reduced
form coefficients than structural ones, so there was no unique solution
for the latter.
So how can we estimate the supply function when it is over-identified?
Since we have no basis for deciding between the exogenous variables in
choosing our instrument why not choose both of them? Or, more precisely,
a linear combination of them. This sounds promising, but what linear
combination should we take (what weight should we give each variable)?
444 Econometrics for developing countries
Why not use the coefficient given by regressing the variable we wish to
proxy on the instruments (the instruments being all the exogenous vari-
ables in the model)? That is, proxy the endogenous regressors by their
fitted values from the estimates of their reduced forms. This method is
perfectly acceptable, and it will yield identical results to the closely related
method of two-stage least squares. But to carry out 2SLS, rather than
using the fitted values of the endogenous variables in the formula for the
IV estimator, we use the fitted values as the regressors in the original
structural equation we are estimating rather than the actual values of these
variables.
Thus, the procedure is called two-stage least squares since it involves
doing two sets of regressions. First we regress the endogenous regressors
from the structural equation we wish to estimate on their reduced forms.
Then we estimate the structural equation by OLS, but replacing the
endogenous regressors by the fitted values given by the first stage. 5 We
now apply the technique to the supply and demand model.
The first stage is to regress the endogenous regressor on its reduced
form; this gives:
Pt = -1.14 + 1.19Y1 - 0.03A 1 (14.26)
which is used to calculate the values of P which are then used to esti-
mate the second stage-equation:
Q/ = 8.36 + 0.20.P,
(39.69) (5.38) (14.27)
The standard errors given by this second regression are not the appro-
priate ones. These are, however, not difficult to calculate, and are given
by statistics and econometrics packages. 6 Also it is not necessary to carry
out the two stages yourself in practice as this will be done by the package
(though it is useful for understanding the technique to do it yourself a
few times and, as we shall see, it can yield useful information). The esti-
mated equation given by selecting 2SLS is:
A A

Q/ = 8.36 + 0.20P1 R2 = 0.04


(30.57) (4.14) (14.28)
Whilst the coefficients are the same as those reported in equation (14.27),
we can see that the correct standard errors are somewhat different to
(larger than) those given by the second-stage regression.
The results from 2SLS are not the same as those we obtained by IV and
ILS earlier in the chapter - there the supply elasticity was 4.82, now its is
4.43. The difference arises because the 2SLS estimates have been obtained
from a different model; that is, one including advertising in the demand
equation. Our estimate of the supply elasticity has therefore been shown
to be reasonably, but not totally robust, to model specification: what we
decide to leave out of the demand equation affects the parameter estimates
Estimating simultaneous equation models 445
of the supply equation. For a limited information technique it is only
which other variables are included in the model that affects the results,
not the precise specification of the other equations; we shall see that this
is not the case for full information techniques, for which misspecification
bias can also arise for all equations because of misspecification of a single
equation.
If we had been estimating the same model, then it would have been
the case that 2SLS would be equivalent to IV - so that 2SLS is also consis-
tent but not unbiased: yet again small-sample bias might be a serious
problem. The technique also assumes that the errors from the different
equations in the system are not correlated with one another - we return
to this in section 14.6. Because of the equivalence between 2SLS and IV,
it is common just to apply 2SLS to any equation which is identified
(whether just or over-identified). 7 Where the equation is just identified
2SLS will yield identical results to IV and therefore also an identical result
to using ILS (why?).
The R 2 from the first-stage regression (not available if we execute the
whole procedure as a computer package option) can give useful infor-
mation on the results. If this R 2 is low then the endogenous regressor is
poorly explained by the variables in its reduced form - thus the fitted
value from this first-stage regression will be a poor proxy for the actual
value and the second-stage estimates will not be worth much. On the other
hand, if the R 2 from the first stage is high then the fitted and actual values
of the endogenous regressor will be very close, so that the 2SLS and OLS
results will be very similar. A high R 2 is quite likely to occur in models
with a large number of exogenous variables. We showed in Chapter 13
that we can use Hausman's specification test to determine whether or not
the difference between the results from OLS and simultaneous techniques
is significant or not. 8
There are two possible definitions of the R 2 far 2SLS: (a) the square
of the correlation coefficient between the dependent variable and its fitted
value; (b) one minus the ratio of the squared residuals to the total sum
of squares. The two are not the same - for our example above we use
the second method, the first method yields an R 2 of 0.89 - and different
packages report different versions. The resulting R 2 from the second defi-
nition may be higher or lower than that from OLS and, like that from IV
estimation, can actually be negative: if it is negative (or very low) then
(especially if the R 2 from OLS estimation of the structural equation is
high) there is a problem in model specification.

14.5 ESTIMATING THE CONSUMPTION FUNCTION


IN A SIMULTANEOUS SYSTEM
Here we present a further example using the model and data for the
Indonesian consumption function (data file INDONA). The model is:
446 Econometrics for developing countries
et == 131 + l32Yr + et (14.29)
yt == et + lt + Xt - Mt (14.30)
where 1 is investment, X exports and M imports of goods and services.
We will first estímate the parameters of the consumption function using
IV and ILS, and show their equivalence. If J, X and M are each treated
as a separate exogenous variable than it can be shown that the consump-
tion function is over-identified, and we cannot apply either IV or ILS
(why not?). Since they are all three exogenous and do not appear in the
behaviourial equation to be estimated, we can define a new variable IXM
== 1 +X - M, and so rewrite equation (14.29) as:
Y== e+ IXM (14.31)
The consumption function is now just identified. Estimating it using IV
with IXM as the instrument yields (t-statistics in parentheses):
et == 1,388.83 + 0.65Yt R 2 == 0.99
(0.91) (42.28) (14.32)
which compares with the OLS estímate of: 9
et== 24.71 + 0.67PYt R2 == 0.99
(0.02) (45.43) DW == 0.33 (14.33)
As expected, the OLS estímate has an upward bias. 10 There is also a bias
in the estímate of the intercept: although the IV estimator (like that
obtained by OLS) is insignificantly different from zero, which conforms
with our theoretical expectation.
To estímate by ILS we need to calculate the reduced form for the model.
This is:
y= ___A_+ _1_ IXM + _e_t- (14.34)
l-f32 l-f32 t l-f32

(14.35)

Which we estímate as:


Yi == 1T1 + 7TzfXMt + w1,r (14.36)
et== 1T3 + 1TiXMt + wz,t (14.37)
where:

(14.38)
Estimating simultaneous equation models 447
There are only two structural coefficients to be estimated (13 1 and ¡3 2) but
four reduced-form estimates. It might appear that we have too many equa-
tions, so that the consumption function is over-identified. This is not the
case. Applying the order condition shows (subject to the rank condition
being satisfied) that the consumption function is just identified. When an
equation is just identified but there appear to be multiple solutions (as
here ), then, in fact, the solutions arrived at by the different channels will
turn out to be the same. You should be able to satisfy yourself that this
is the case by estimating the 7r from equations (14.36) and (14.37) and
substituting the results in to equation (14.38).
The reduced form for income is:
Yt = 4,011 + 2.888/XMt (14.39)
which gives:
/\

b ILS = 7r2 - 1 = 2.888 - 1 =


2 ,;2 2.888 o. 65 (14.40)

b[LS = (1 - b2) TI¡ = 1,388

The same answers are arrived at by estimating the reduced form for
consumption:

et = 4,011 + l.888IXM 1
(14.41)
so:
/\

b/LS = ~ = 1.888 - 1 = 0.65 (14.42)


7r4 +1 1.888

b[LS = (1 - b2) TI3 = 1,388


thus showing that it does not matter which reduced form is used to derive
the results. 11 It can also be seen that, as expected, the ILS estimates are
the same as those arrived at by IV.
We can also estimate the model's parameters by two-stage least squares,
in which case there is no need to ensure that the consumption function
is just identified. Instead /, X and M are treated as three separate exoge-
nous variables. 12 The two stages are therefore to estimate the reduced
form for Y, by regressing it on these three variables (anda constant), and
then to regress the consumption function, replacing Y by Y from the first
stage. Both stages have been done in a single step by an econometrics
package to get:
et = 232.89 + 0.67Yt R 2 = 0.99 (14.43)
(0.16) (45.09)
448 Econometrics far developing countries
This gives a slightly better fit than IV using IXM as a single instrument.
This follows since the fitted value from regressing Y on /, X and M sepa-
rately is sure to be a better proxy for Y than is IXM (why?).

Exercise 14.5
Use the Indonesian data (data file INDONA) to estimate the model:
Yt = et + It + xt - Mt (14.44)
et= f31 + f32 yt + El,t (14.45)
Mt = ª1 + ª2Yt + Ez,t (14.46)

Exercise 14.6
Repeat the procedures followed in the preceding section using the national
accounts data for Sri Lanka (data file SRINA) or the country of your
choice.

14.6 FULL INFORMATION ESTIMATION TECHNIQUES


The techniques considered so far are limited information: they estimate
one equation at a time and only use information pertaining to that equa-
tion. Whilst these techniques are consistent, they are not as efficient as
full information (or systems) estimation, in which all model parameters
are estimated in one go. The inefficiency of limited information estima-
tion arises since the method <loes not exploit two possible pieces of
information: (a) cross-correlation between error terms in different equa-
tions; and (b) that the predetermined variable omitted from the equation
being estimated may also be omitted from other equations in the system.
Seemingly unrelated regression (SUR) takes account of just the first of
these, and three-stage least squares (3SLS) takes account of both. These
techniques use generalised least squares (GLS) estimation, which is
beyond the scope of our text. Hence we focus on an intuitive explana-
tion, so that readers will understand what these methods are and why
they are being used, if they should come across them.

Seemingly unrelated regressions


In our discussion of recursive systems in section 14.2, we stated that OLS
estimation of each equation is consistent only if there is no cross-corre-
lation between the different error terms. Where such cross-correlation
exists then an alternative technique should be used. In fact, if there is
cross-correlation then, even if the regressions appear unrelated to one
another, it is preferable to utilise this information and hence improve the
Estimating simultaneous equation models 449
efficiency of the estimates. A common example of appropriate applica-
tion of SUR is the demand equations for related products:
Ql,t = ª1 + ª2P1,1 + E1,1 (14.47)
Q2,t = 131 + l32P2,t + E2,1 (14.48)
where P and Q are price and quantity respectively and 1 and 2 are two
commodities (e.g. tea and coffee). To apply SUR estimation the samples
for the two regressions must be the same (e.g. same period of time or
cross-section of households).
SUR estimation combines the two equations into a single equation for
estimation purposes, and applies generalised least squares to this equa-
tion. Here we state without proof the expressions for the two slope
coefficien ts:
suR _ [u; LP1 ql - <T12LP1 q2]ui LP; + [ui LP2q2 - <T12LP2qiJu12LP1P2
ª2 - 2 2"" 2"" 2 2 (""
0"1 O"z ,:.,pi ,:.,pz - 0"12 ,:.,P1P2
)2
(14.49)

bsuR _ [u; LP1 q¡ - <T12LP1 q2]<T12LPiP2 + [ui LP2q2 - <T12LP2 qi]af LP~
2 2
- uiu;LPiLP; - u1 2(LP1P2)2
(14.50)
Examination of equations (14.49) and (14.50) shows that if there is no
cross-equation error correlation then the estimators are equivalent to
OLS. SUR estimators are also equivalent to OLS when the two regres-
sors (P1 and P 2) are identical.
In arder to apply these equations, estimates of the error variance and
covariance are required, which are calculated from the residuals:

(14.51)

Table 14.1 Estimates for two-commodity supply and demand model


lntercept Slope R2
Ordinary least squares
Ql 5.78 -1.44 0.90
(19.6) (-19.2)
Q2 3.06 -0.52 0.73
(16.3) (-10.8)
Seemingly unrelated
regressions estimation
Ql 5.87 -1.47 0.89
(28.9) (-28.2)
Q2 3.04 -0.51 0.73
(23.4) (-15.5)
450 Econometrics far developing countries
Table 14.2 Cross-producís for SUR estimation
e¡ e2 P1 P2 q¡ q2
e1 0.247 0.090
ez 0.066
P1 1.007 -0.074 -1.456 0.056
P2 0.677 0.097 -0.351
q¡ 2.351 0.014
qz 0.248

where n is the number of observations and e1 and e2 are the residuals


from OLS estimation of the first and second equations.
To illustrate the technique we apply the data from data file SURE,
which contains data simulated using the model in equations (14.47) and
(14.48), with the values a 2 = -1.5 and [3 2 = -0.5. Table 14.1 shows the
results obtained by OLS and SURE, and Table 14.2 the cross-products
required to estimate the slope coefficients. 13 The results show not much
difference in the coefficient estimates between the two techniques (the
SUR estimates are slightly closer to the actual population values). But
there is a substantial difference in the t-statistics: as we would expect,
since SUR estimation is the more efficient technique, it produces lower
standard errors and correspondingly larger t-statistics.

Exercise 14.7
Using the data in SOCECON estimate the regressions of the birth rate
on the square root of infant mortality, and of life expectancy on logged
income per capita using (a) OLS and (b) SUR. Comment on your results.

Three-stage least squares


SUR estimation is a two-stage technique in which the residuals are
calculated from OLS regression in order to apply generalised least
squares. Three-stage least squares (3SLS) is similar to SUR, except that
the residuals are obtained from the second-stage regressions of TSLS.
Three-stage least squares may be applied to simultaneous models and, like
SUR, results in more efficient estimates (usually around 5 per cent more
efficient that those produced by TSLS). If there are cross-equation restric-
tions (i.e. the parameters in one equation are related to those in another)
then systems estimation must be applied and 3SLS is the approach to
be adopted.
To summarise, the three stages of 3SLS are:
1 Estimate the reduced form equations by OLS.
2 Substitute the fitted values from stage 1 into the structural equations
and estimate by OLS (i.e. the second stage regression of TSLS).
Estimating simultaneous equation models 451
Table 14.3 Estimates of simple Keynesian model for Indonesia
lntercept Slope R2
Ordinary least squares
Consumption 24.71 0.067 0.99
(0.02) (45.4)
Imports -2862.7 0.24 0.90
(-1.67) (14.1)
Two-stage least squares
Consumption -23.5 0.67 0.99
(-0.2) (45.2)
Imports -3504.5 0.25 0.90
(-2.0) (14.4)
Three-stage least squares
Consumption -23.54 0.67 0.99
(-0.2) (47.1)
Imports -3504.5 0.25 0.90
(-2.1) (15.0)
Source: Calculated from data file INDONA
Note: t-statistics in parentheses

3 Calculate the residuals from stage 2 to obtain an estimate of the error


variance-covariance matrix and so apply generalised least squares.
Modern econometrics packages will produce 3SLS estimates. Table 14.3
reports the results for the simple Keynesian model for Indonesia with
consumption and import functions (see exercise 14.5). The coefficient esti-
mates do not differ greatly between 3SLS and TSLS (though sometimes
they do with OLS because the latter is biased). But the 3SLS t-statistics
are all larger than those obtained from TSLS.

Exercise 14.8
Replicate Table 14.3 using the INDONA data. Reproduce the table also
using the data in SRINA or national accounts data for the country of your
choice. Comment on your results.

14.7 SUMMARY OF MAIN POINTS


Simultaneous models may be estimated by a range of techniques. Indirect
least squares is intuitively appealing, but too cumbersome to use for all
but the simplest of models. Two-stage least squares is commonly used,
though the equivalent technique of instrumental variables may be used
for just identified equations.
ILS, IV and TSLS are single equation estimation techniques, as each
equation is estimated in turn. Full information, or systems estimation,
produces estimates for all model parameters in one go. Systems estima-
tion can take account of the fact that there may be cross-correlation
452 Econometrics for developing countries
between the error terms of different equations and so it gives more effi-
cient estimates. Two words of warning are in arder. With systems
estimation misspecification of a single equation will bias the parameter
estimates far all equations. Second, the problem of non-stationarity is
more serious than that of simultaneity bias - and a set of cointegrating
equations will not suffer from simultaneity bias even if the equations
constitute a simultaneous model.

NOTES
1 Standard errors are not reported as there is no simple way of obtaining them
from the reduced-form standard errors (which were not reported for the same
reason - their significance does not necessarily imply the significance of the
structural coefficients). The inability to judge significance from ILS estimation
is an additional reason against its use to that given later in the text.
2 The reason being that an attempt to prove unbiasedness runs into the need
to take expectations of a ratio of random variables. This is not possible, and
we must resort to probability limits.
3 Many packages, for example Microfit, treat the intercept as a variable. Since
to estimate an equation by IV we need as many instruments as we have vari-
ables, estimation of the supply schedule requires two instruments. But the
intercept (or any other exogenous variable appearing in the equation) simply
acts as its own instrument.
4 We gave the formula for the slope coefficient (but not the intercept and stan-
dard errors) in Chapter 13.
5 We can actually replace ali the endogenous variables - including that on the
left-hand side of the equation - by their fitted values, as using fitted rather
than actual values on the left-hand side does not affect the parameter esti-
mates. (It does change the standard errors, but these are, as we shall see, the
wrong ones anyway.)
6 To get the correct standard errors from the ones given by the second stage it
is necessary to multiply by a correction factor. This correction factor is the
ratio of the estimated variance of the error in the structural equation being
estimated to the variance of the error in the second-stage regression (the struc-
tural equation with endogenous regressors replaced by their fitted values from
the first stage regression). See Gujarati (1988: 620-1) or Maddala (1988:
311-13) for a derivation.
7 In which case, as stated in Chapter 13, it is unnecessary to check the order
condition, since the rank condition is necessary and sufficient.
8 In Chapter 13 we used the IV method to provide a consistent estimator when
the regressor is endogenous. It is possible to use 2SLS instead.
9 Note that the DW is substantially less than the R 2 , suggesting that there is
likely to be spurious correlation here and that cointegration analysis might be
appropriate. As stated in Chapter 11, this latter approach should be adopted
- using simultaneous techniques does not eliminate the inconsistency that
results from regressing series that are I(O) on one another.
10 However, the bias is very small. Sorne suggest simultaneity is never too great
a problem - and almost certainly less than the measurement error in the data.
We should also note with concern that R2 > DW, suggesting that the regres-
sion may well be spurious. It is more important to test for stationarity, and
proceed accordingly if the variables prove non-stationary, than to worry about
simultaneous estimation.
11 lf different results are obtained this tells you there is a problem in the data.
This is why we could not use the textbook example of Y = C + !, since there
is no economy which fits this model. lf data are used for which this identity
does not hold then different results will be obtained from the two reduced
forms. The Y here is GDP at market prices: GNP might be a more appro-
priate measure to use in the consumption function, in which case net factor
payments from abroad should be added as an additional exogenous term in
the identity.
12 Endogenising imports through an import function, which may appear an
obvious improvement to the model, is left as an exercise.
13 We have not given the formulae for the intercept and standard errors. These
estimates were obtained from TSP.
This page intentionally left blank
Appendix A
The data sets used in this book

This appendix gives a brief description of each of the sets available on


the data diskette and used in the examples and exercises of this book.
For your convenience, the data sets are listed in alphabetical arder of file
name. For each data set, a brief description is given of the nature of the
data, its sources, sample size, and a list and definition of the variables.
The variables are ordered by columns in each case. For reasons of confi-
dentiality, however, the exact reference to sorne of the data sets which
contain survey data for individual cases is kept vague to avoid identifica-
tion of case materials.
The arder of the variables in each of the data files in ASCII is the same
as they appear in the following description of the data files. Since DOS
allows only eight characters for the first name of a file, the names KRISH-
NAJI and FERTILITY (mentioned in the book) appear as KRISHNAJ
and FERTILIT for the data files. The Lotus files include the variable
names, but these are excluded in the ASCII files.

AIDSAV (aid and savings)


Cross-country data for 66 developing countries for 1987 taken from the
World Development Report. The variables are:
1 observation numbers (OBS);
2 the savings rate (S/Y);
3 aid as a per cent of GDP (A/Y); and
4 income per capita in US$ (Y).

BELPOP (Belize's population)


Population data for Belize for the years 1970-92 (from the World Bank's
World Tables). The file contains the following variables:
1 YEAR (1970-92); and
2 population (POP) in millions.
456 Econometrics for developing countries
BIRTH (the birth rate and its determinants)
The file BIRTH contains data (taken from the the World Bank's World
Tables) for 128 countries in 1985 on the following variables:
1 the crude birth rate (BIRTH): the ratio of the number of births in ayear
over the population in the mid-year, expressed per 1,000 population;
2 the gross national income per capita ( GNPCAP);
3 infant mortality (INFMORT): the number of deaths of infants under
age one in ayear over the number of live births in that year, expressed
per 1,000 live births.

CHINA (household size and income data)


The file contains data sub-sampled from the World Fertility Survey (1975),
for 197 rural households, Hebei Province, China. The file contains the
following variables:
1 household size (SIZE);
2 annual household income (INCOME) in US$;
3 number of children (TPARITY);
4 number of couples (COUPLES);
5 per capita household income (PI).

COINCON (a cointegrating consumption function)


Simulated data to estímate a cointegrating consumption function. The data
span the period 1960-93 and are in logs. The variables are:
1 YEAR;
2 logged consumption (LC); and
3 logged income (LY).

CRCON (Costa Rican consumption function)


Logged data to estimate a consumption function for Costa Rica for
the period 1963-92. Sources are the World Bank's World Tables and
International Finance Statistics. The file contains:
1 YEAR;
2 logged consumption (LC); and
3 logged income (LY).

FERTILIT (the determinants of fertility data)


This file contains comparative cross-section data for 64 countries on
fertility and its determinants as given by the following variables:
1 GNP per capita 1980 ( GNP);
Appendix A: Data sets 457
2 the total fertility rate (TFR), 1980-85: the average number of children
born to a woman, using age-specific fertility rates for a given year;
3 child mortality (CM): the number of deaths of children under age five
in a year per 1,000 live births;
4 female literacy rate (FL), expressed as percentage;
5 an index of family planning effort (FP).
With the exception of GNP per capita (taken from the World Bank's
World Tables), all other data are taken from J. A. Ross et al., Planning
and Child Survival: 100 Developing Countries, Centre for Population and
Family Health, Columbia University, 1988.

INDFOOD (Indian households food data)


This file contains data for a sample of 217 Indian rural households (1994)
with the following variables;
1 household expenditures on food (F);
2 total expenditures of the household (T).

INDIA (industrial worker data, southern India)


This file contains data for a sample of 261 workers collected in a survey
of worker households in 1990 in an industrial town in southern India. The
file contains the following variables:
1 weekly wage income (WJ);
2 sex of worker (SEX);
3 dummy variable to indicate sex of worker (DSEX);
4 education of worker (EDU): none, primary, secondary, higher;
5 dummies for education: DE2 (up to primary = 1; O otherwise); DE3
(up to secondary =1; Ootherwise); DE4 (higher education = 1; O other-
wise);
6 permanent or temporary job (PT);
7 dummy for permanent/temporary job (DPT) (DPT =1 if permanent;
O otherwise );
8 age of worker in years (AGE).

INDONA (lndonesian national accounts data)


Constant price national accounts for Indonesia for the period 1968-92 in
billions of Rupiah. (Taken from the World Bank's World Tables and the
IMF's International Finance Statistics.) The variables are:
1 YEAR;
2 GDP (Y);
3 total consumption ( C);
458 Econometrics far developing countries
4 investment (I);
5 exports (X); and
6 imports (M).

KRISHNAJ (food prices and manufacturing demand)


The data for this file are taken from Krishnaji (1992: 102-3, tables 3 and
4). The data are time series for the period 1960-1 to 1980-1, and include
the following variables:
1 disposable income (in Rupees) per capita in current prices (Y);
2 per cap ita expenditures (in Rs) on manufactured consumer goods at
constant 1970-1 prices (M);
3 price index (1970-1 = 100) of cereals and cereal substitutes (P1);
4 price index (1970-1 = 100) of other foods and beverages (P 01);
5 price index (1970-1 = 100) of manufactured consumer goods (P m);
6 price index (1970-1 = 100) of services (Ps).

LEACCESS (determinants of life expectancy)


Data for 87 countries in the late 1980s taken from the UNDP's Human
Development Report. The variables are:
1 country name (not included in ASCII file);
2 life expectancy at birth (LE);
3 income per capita in US$ (Y); and
4 access to health care (ACCESS).

MALTA (Maltese exports demand and supply)


These data, from Gatt (1995), cover the period 1963-89. The variables are:
1 YEAR;
2 exports in US$ (X);
3 a measure of world demand (Y);
4 the export price (P);
5 the world price (PW);
6 the nominal exchange rate (E);
7 consumer price index (CPI); and
8 capital investment (I).

MAPUTO (the demand for and recruitment of


casual labour, Maputo harbour)
This file contains data on the daily demand for and recruitment of casual
labourers in Maputo harbour for the period from March 1980 to June
Appendix A: Data sets 459
1981, a sample of 485 consecutive days (with the exception of a few holi-
days such as Christmas and New Year). The file contains the following
variables:
1 the demand for labour on the day shift (DEMD), expressed in number
of workers;
2 the demand for labour on the night shift (DEMN), expressed in
number of workers;
3 the total demand for labour (DEM), both day and night shift;
4 the recruitment of labour on the day shift (RECD), expressed in
number of workers;
5 the recruitment of labour on the night shift (RECN), expressed in
number of workers;
6 the total recruitment of labour (REC), both day and night shift;
7 the trend variable (t), indicating successive working days.

PAKEXP (Pakistani export function)


Data for the period 1966--92 from International Finance Statistics (and the
Direction of Trade Statistics to determine weights to calculate the RER).
The variables are:
1 YEAR;
2 real exports (X);
3 world demand (WD); and
4 the real exchange rate (RER).

PERU (data base for Pero)


A database for the period 1967-93 constructed from the World Bank's
World Tables and International Financia! Statistics. The variables are
YEAR and:
1 population (POP);
2 nominal GDP (YC);
3 real GDP (Y);
4 real consumption ( C);
5 real exports (X);
6 real imports (M);
7 real manufacturing value added (MVA);
8 the externa! terms of trade (TOT);
9 life expectancy at birth (LE);
10 infant mortality rate (IMR);
11 externa! current account deficit as a percentage of GDP (CA);
12 nominal interest rate (R);
13 real interest rate (r); and
14 consumer price index ( CPI).
460 Econometrics for developing countries
PRODFUN (production function data)
Data from a business survey in a developing country reporting the data
far two manufacturing sectors and covering 109 firms. The fallowing vari-
ables are included:
1 SECTOR (4 or 9 corresponding to the sector number);
2 logged output (Q)
3 logged capital stock (K); and
4 logged employment (L).

SIMSIM (supply and demand model)


Data simulated (40 observations) from the model in equations 13.1 to
13.3. The variables are:
1 OBS;
2 quantity (Q);
3 price (P); and
4 income (Y).
5 advertising (ADV)

SOCECON (world socioeconomic data)


The file SOCECON contains a selection of socioeconomic data, taken from
various sources as indicated below, far a total of 125 countries with 1990 as
the reference year. The sample size far different exercises or examples
will vary due to missing observations far sorne of the variables involved.
Data are from the World Bank's World Tables and Social Indicators of
Development and the UNDP's Human Development Report. The data set
contains the fallowing variables:
1 GNP per capita, 1990 (data from the World Bank's World Tables);
2 LEX, 1990: life expectancy at birth (World Development Report 1992);
3 LEXPF, 1990: life expectancy female (World Development Report
1992);
4 LEXPM, 1990: life expectancy male (World Development Report
1992);
5 ADULIT, 1990: adult literacy rate (per cent) (Human Development
Report 1993);
6 HDI, 1990: human development index (Human Development Report
1993);
7 FERTR, 1990: total fertility rate (World Development Report 1992);
8 BIRTHR, 1990: crude birth rate per 1,000 population (World
Development Report 1992);
9 POP, 1990: population (World Bank's World Tables);
10 POPGTH, 1990: annual population growth rate (per cent) World Social
Indicators data base);
Appendix A: Data sets 461
11 CHILDMF, 1990: female under 5 mortality rate per 1,000 live births
(World Development Report 1992);
12 CHILDMM, 1990: male under 5 mortality rate per 1,000 live births
(World Development Report 1992);
13 INFMOR, 1990: infant mortality rate per 1,000 live births (World
Development Report 1992);
14 URBANPOP, 1990: urban population as a percentage of total popu-
lation 1991 (World Bank's World Tables);
15 ENERGCPC, 1990: energy consumption per capita (World Bank's
World Tables);
16 PPPGDP: per capita gross domestic product in purchasing power party
dollars;
17 HEALTHAC: per cent of population with access to health services;
18 WATERAC: per cent of population with access to clean water;
19 SANITAT: per cent of population with access to sanitation facilities;
20 MVA: manufacturing value added (as a per cent of GDP);
21 COUNTRY, country long name (not included in ASCII file);
22 CID, country short name (not included in ASCII file).

SRINA (Sri Lanka macroeconomic data)


A macroeconomic data set for the years 1967-93. All variables are in
constant prices. (From the World Bank's World Tables and G. Pfefferman
and A. Madarassay (various years) Trends in Prívate lnvestment in
Developing Countries, New York: World Bank, for the public/private
investment split.) All aggregates are in constant prices. The variables are:
1 YEAR;
2 GDP;
3 exports (X);
4 imports (M);
5 private consumption ( CP);
6 government consumption (Cg);
7 total consumption ( CN);
8 investment (/);
9 private investment (IP);
10 government investment (/g);
11 real interest rate (r).

SURE (data for seemingly unrelated regressions)


Data simulated from equations (14.47) and (14.48). The variables are:
1 YEAR;
2 quantity for good 1 (Q 1);
3 quantity for good 2 (Q 2);
4 price for good 1 (P1); and
5 price for good 2 (P2 ).
462 Econometrics for developing countries
TANMON (money and prices in Tanzania)
Data from the World Bank's World Tables and IMF's International Finance
Statistics for the years 1966-92, covering:
1 YEAR;
2 consumer price index ( CPI);
3 money supply (M2); and
4 real export growth (XGRO).

TANZANIA (government expenditure and revenue data)


The following data on government finance are taken from the Tanzanian
Economic Surveys for the period 1970-90. All data series are deflated by
the consumer price index:
1 government recurrent revenue (REV);
2 government recurrent expenditures (REXP);
3 government development expenditures (DEXP);
4 year (1): 1970-90.

TOT ( developing countries' terms of trade)


The data are the ratio of price indices for non-fuel primary commodities
and manufactured goods for the period 1950-86 taken from Grilli and
Yang (1988). The file contains:
1 YEAR; and
2 terms of trade (TOI).

TPEASANT (farm size and household size data, rural Tanzania)


This file contains grouped data (eight classes) based on a sample of 600
peasant households as published by Collier et al. (1986: 50, Table 3.12).
The data are grouped by classes of farm sizes. The file has the following
variables:
1 mean farm size (F) in acres;
2 mean household size (H): number of persons;
3 per cent of households in each group (percent).
Appendix B Statistical tables

Note: Tables B.1-B.6 are reproduced, with permission, from D. Gujarati,


Basic Econometrics, 2nd edn, New York: McGraw-Hill, 1988. The orig-
inal sources are noted on each table.
464 Econometrics far developing countries
Table B.1 Areas under the standardised normal distribution

Example
Pr (O ,,;; z ,,;; 1.9607 = 0.4750
Pr (z ~ 1.96) = 0.5 - 0.4750 = 0.025

o 1.96

z 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
O.O 0.0000 0.0040 0.0080 0.0120 0.0160 0.0199 0.0239 0.0279 0.0319 0.0359
0.1 0.0398 0.0438 0.0478 0.0517 0.0557 0.0596 0.0636 0.0675 0.0714 0.0753
0.2 0.0793 0.0832 0.0871 0.0910 0.0948 0.0987 0.1026 0.1064 0.1103 0.1141
0.3 0.1179 0.1217 0.1255 0.1293 0.1331 0.1368 0.1406 0.1443 0.1480 0.1517
0.4 0.1554 0.1591 0.1628 0.1664 0.1700 0.1736 0.1772 0.1808 0.1844 0.1879
0.5 0.1915 0.1950 0.1985 0.2019 0.2054 0.2088 0.2123 0.2157 0.2190 0.2224

0.6 0.2257 0.2291 0.2324 0.2357 0.2389 0.2422 0.2454 0.2486 0.2517 0.2549
0.7 0.2580 0.2611 0.2642 0.2673 0.2704 0.2734 0.2764 0.2794 0.2823 0.2852
0.8 0.2881 0.2910 0.2939 0.2967 0.2995 0.3023 0.3051 0.3078 0.3106 0.3133
0.9 0.3159 0.3186 0.3212 0.3238 0.3264 0.3289 0.3315 0.3340 0.3365 0.3389
1.0 0.3413 0.3438 0.3461 0.3485 0.3508 0.3531 0.3554 0.3577 0.3599 0.3621

1.1 0.3643 0.3665 0.3686 0.3708 0.3729 0.3749 0.3770 0.3790 0.3810 0.3830
1.2 0.3849 0.3869 0.3888 0.3907 0.3925 0.3944 0.3962 0.3980 0.3997 0.4015
1.3 0.4032 0.4049 0.4066 0.4082 0.4099 0.4115 0.4131 0.4 147 0.4162 0.4177
1.4 0.4192 0.4207 0.4222 0.4236 0.4251 0.4265 0.4279 0.4292 0.4306 0.4319
1.5 0.4332 0.4345 0.4357 0.4370 0.4382 0.4394 0.4406 0.4418 0.4429 0.4441

1.6 0.4452 0.4463 0.4474 0.4484 0.4495 0.4505 0.4515 0.4525 0.4535 0.4545
1.7 0.4554 0.4564 0.4573 0.4582 0.4591 0.4599 0.4608 0.4616 0.4625 0.4633
1.8 0.4641 0.4649 0.4656 0.4664 0.4671 0.4678 0.4686 0.4693 0.4699 0.4706
1.9 0.4713 0.4719 0.4726 0.4732 0.4738 0.4744 0.4750 0.4756 0.4761 0.4767
2.0 0.4772 0.4778 0.4783 0.4788 0.4793 0.4798 0.4803 0.4808 0.4812 0.4817

2.1 0.4821 0.4826 0.4830 0.4834 0.4838 0.4842 0.4846 0.4850 0.4854 0.4857
2.2 0.4861 0.4864 0.4868 0.4871 0.4875 0.4878 0.4881 0.4884 0.4887 0.4890
2.3 0.4893 0.4896 0.4901 0.48901 0.4904 0.4906 0.4909 0.4911 0.4913 0.4916
2.4 0.4918 0.4920 0.4922 0.4925 0.4927 0.4929 0.4931 0.4932 0.4934 0.4936
2.5 0.4938 0.4940 0.4941 0.4943 0.4945 0.4946 0.4948 0.4949 0.4951 0.4952

2.6 0.4953 0.4955 0.4956 0.4957 0.4959 0.4960 0.4961 0.4962 0.4963 0.4964
2.7 0.4965 0.4966 0.4967 0.4968 0.4969 0.4970 0.4971 0.4972 0.4973 0.4974
2.8 0.4974 0.4975 0.4976 0.4977 0.4977 0.4978 0.4979 0.4979 0.4980 0.4981
2.9 0.4981 0.4982 0.4982 0.4983 0.4984 0.4984 0.4985 0.4985 0.4986 0.4986
3.0 0.4987 0.4987 0.4987 0.4988 0.4988 0.4989 0.4989 0.4990 0.4990 0.4990
Note: This table gives the area in the right-hand tail distribution (i.e., z ~ O). But since the normal distribution is
symmetrical about z = O , the area in the left-hand tail is the same as the area in the corresponding right-hand tail. For
example, P(-1.96 "°' z "°'O) = 0.4750. Therefore, P(-1.96 ,o; z ,o; 1.96) = 2(0.4750) = 0.95.
Appendix B: Statistical tables 465

:;.:::~:: pob" of fuo < d;,.,Moo~ O.OS

Pr (t > 1.725) = 0.05 for df = 20 ~----- 0 l.725


Pr (!ti > 1.725) = 0.10

~
0.025 O.JO 0.05 0.025 O.OJ 0.005 O.OOJ
f 0.50 0.20 O.JO 0.05 0.02 O.OJO 0.002

1 1.000 3.078 6.314 12.706 31.821 63.657 318.31


2 0.816 1.886 2.920 4.303 6.965 9.925 22.327
3 0.765 1.638 2.353 3.182 4.541 5.841 10.214
4 0.741 1.533 2.132 2.776 3.747 4.604 7.173
5 0.727 1.476 2.015 2.571 3.365 4.032 5.893
6 0.718 1.440 1.943 2.447 3.143 3.707 5.208
7 0.711 1.415 1.895 2.365 2.998 3.499 4.785
8 0.706 1.397 1.860 2.306 2.896 3.350 4.501
9 0.703 1.383 1.833 2.262 2.821 3.250 4.297
10 0.700 1.372 1.812 2.228 2.764 3.169 4.144
11 0.698 1.363 1.796 2.201 2.718 3.106 4.025
12 0.695 1.356 1.782 2.179 2.681 3.055 3.930
13 0.694 1.350 1.771 2.160 2.650 3.012 3.852
14 0.692 1.345 1.761 2.145 2.624 2.977 3.787
15 0.691 1.341 1.753 2.131 2.602 2.947 3.733
16 0.690 1.337 1.746 2.120 2.583 2.921 3.686
17 0.689 1.333 1.740 2.110 2.567 2.898 3.646
18 0.688 1.330 1.734 2.101 2.552 2.878 3.610
19 0.688 1.328 1.729 2.093 2.539 2.861 3.579

20 0.687 1.325 1.725 2.086 2.528 2.845 3.552


21 0.686 1.323 1.721 2.080 2.518 2.831 3.527
22 0.686 1.321 1.717 2.074 2.508 2.819 3.505
23 0.685 1.319 1.714 2.069 2.500 2.807 3.485
24 0.685 1.318 1.711 2.064 2.492 2.797 3.467
25 0.684 1.316 1.708 2.060 2.485 2.787 3.450
26 0.684 1.315 1.706 2.056 2.479 2.779 3.435
27 0.684 1.314 1.703 2.052 2.473 2.771 3.421
28 0.683 1.313 1.701 2.048 2.467 2.763 3.408
29 0.683 1.311 1.699 2.045 2.462 2.756 3.396
30 0.683 1.310 1.697 2.042 2.457 2.750 3.385
40 0.681 1.303 1.684 2.021 2.423 2.704 3.307
60 0.679 1.296 1.671 2.000 2.390 2.660 3.232
120 0.677 1.289 1.658 1.980 2.358 2.167 3.160
00 0.674 1.282 1.645 1.960 2.326 2.576 3.090
Note: The smaller probability shown at the head of each column is the area in one tail; the larger probability is the
area in both tails.
Source: From E.S. Pearson and H.O. Hartley (eds), Biometrika Tables far Statisticians, 3rd edn, New York: Cambridge
University Press, 1966.
466 Econometrics far developing countries
Table B.3 Upper percentage points of the F distribution

Example
Pr (F > 1.59) = 0.25
Pr (F > 2.42) = 0.10 for df N 1 = 10
Pr (F > 3.14) = 0.05 and N 2 = 9
Pr (F > 5.26) = 0.01

a¡ ¡ar
denom- df far numerator N 1
inator
N,
Pr 1 2 3 4 5 6 7 8 9 10 11 12
0.25 5.83 7.50 8.20 8.58 8.82 8.98 9.10 9.19 9.26 9.32 9.36 9.41
1 0.10 39.9 49.5 53.6 55.8 57.2 58.2 58.9 59.4 59.9 60.2 60.5 60.7
0.05 161 200 216 225 230 234 237 239 241 242 243 244

0.25 2.57 3.00 3.15 3.23 3.28 3.31 3.34 3.35 3.37 3.38 3.39 3.39
2 0.10 8.53 9.00 9.16 9.24 9.29 9.33 9.35 9.37 9.38 9.39 9.40 9.41
0.05 18.5 19.0 19.2 19.2 19.3 19.3 19.4 19.4 19.4 19.4 19.4 19.4
0.01 98.5 99.0 99.2 99.2 99.3 99.3 99.4 99.4 99.4 99.4 99.4 99.4

0.25 2.02 2.28 2.36 2.39 2.41 2.42 2.43 2.44 2.44 2.44 2.45 2.45
3 0.10 5.54 5.46 5.39 5.34 5.31 5.28 5.27 5.25 5.24 5.23 5.22 5.22
0.05 10.1 9.55 9.28 9.12 9.01 8.94 8.89 8.85 8.81 8.79 8.76 8.74
0.01 34.1 30.8 29.5 28.7 28.2 27.9 27.7 27.5 27.3 27.2 27.1 27.0

0.25 1.81 2.00 2.05 2.06 2.07 2.08 2.08 2.08 2.08 2.08 2.08 2.08
4 0.10 4.54 4.32 4.19 4.11 4.05 4.01 3.98 3.95 3.94 3.92 3.91 3.90
0.05 7.71 6.94 6.59 6.39 6.26 6.16 6.09 6.04 6.00 5.96 5.94 5.91
0.01 21.2 18.0 16.7 16.0 15.5 15.2 15.0 14.8 14.7 14.5 14.4 14.4

0.25 1.69 1.85 1.88 1.89 1.89 1.89 1.89 1.89 1.89 1.89 1.89 1.89
5 0.10 4.06 3.78 3.62 3.52 3.45 3.40 3.37 3.34 3.32 3.30 3.28 3.27
0.05 6.61 5.79 5.41 5.19 5.05 4.95 4.88 4.82 4.77 4.74 4.71 4.68
0.01 16.3 13.3 12.1 11.4 11.0 10.7 10.5 10.3 10.2 10.1 9.96 9.89

0.25 1.62 1.76 1.78 1.79 1.79 1.78 1.78 1.78 1.77 1.77 1.77 1.77
6 0.10 3.78 3.46 3.29 3.18 3.11 3.05 3.01 2.98 2.96 2.94 2.92 2.90
0.05 5.99 5.14 4.76 4.53 4.39 4.28 4.21 4.15 4.10 4.06 4.03 4.00
0.01 12.2 10.9 9.78 9.15 8.75 8.47 8.26 8.10 7.89 7.87 7.79 7.72

0.25 1.57 1.70 1.72 1.72 1.71 1.71 1.70 1.70 1.69 1.69 1.69 1.68
7 0.10 3.59 3.26 3.07 2.96 2.88 2.83 2.78 2.75 2.72 2.70 2.68 2.67
0.05 5.59 4.74 4.35 4.12 3.97 3.87 3.79 3.73 3.68 3.64 3.60 3.57
0.01 12.2 9.55 8.45 7.85 7.46 7.19 6.99 6.84 6.72 6.62 6.54 6.47

0.25 1.54 1.66 1.67 1.66 1.66 1.65 1.64 1.64 1.63 1.63 1.63 1.62
8 0.10 3.46 3.11 2.92 2.81 2.73 2.67 2.62 2.59 2.56 2.54 2.52 2.50
0.05 5.32 4.46 4.07 3.84 3.69 3.58 3.50 3.44 3.39 3.35 3.31 3.28
0.01 11.3 8.65 7.59 7.01 6.63 6.37 6.18 6.03 5.91 5.81 5.73 5.67

0.25 1.51 1.62 1.63 1.63 1.62 1.61 1.60 1.60 1.59 1.59 1.58 1.58
9 0.10 3.36 3.01 2.81 2.69 2.61 2.55 2.51 2.47 2.44 2.42 2.40 2.38
0.05 5.12 4.26 3.86 3.63 3.48 3.37 3.29 3.23 3.18 3.14 3.10 3.07
0.01 10.6 8.02 6.99 6.42 6.06 5.80 5.61 5.47 5.35 5.26 5.18 5.11
Appendix B: Statistical tables 467
Table B.3 - Continued

dffor
df far denominator N 1 denom-
inator
15 20 24 30 40 50 60 100 120 200 500 00 Pr Ni

9.49 9.58 9.63 9.67 9.71 9.74 9.76 9.78 9.80 9.82 9.84 9.85 0.25
61.2 61.7 62.0 62.3 62.5 62.7 62.8 63.0 63.1 63.2 63.3 63.3 0.10 1
246 248 249 250 251 252 252 253 253 254 254 254 0.05

3.41 3.43 3.43 3.44 3.45 3.45 3.46 3.47 3.47 3.48 3.48 3.48 0.25
9.42 9.44 9.45 9.46 9.47 9.47 9.47 9.48 9.48 9.49 9.49 9.49 0.10 2
19.4 19.4 19.5 19.5 19.5 19.5 19.5 19.5 19.5 19.5 19.5 19.5 0.05
99.4 99.4 99.5 99.5 99.5 99.5 99.5 99.5 99.5 99.5 99.5 99.5 0.01

2.46 2.46 2.46 2.47 2.47 2.47 2.47 2.47 2.47 2.47 2.47 2.47 0.25
5.20 5.18 5.18 5.17 5.16 5.15 5.15 5.14 5.14 5.14 5.14 5.13 0.10 3
8.70 8.66 8.64 8.62 8.59 8.58 8.57 8.55 8.55 8.54 8.53 8.53 0.05
26.9 26.7 26.6 26.5 26.4 26.4 26.3 26.2 26.2 26.2 26.1 26.1 0.01

2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 2.08 0.25
3.87 3.84 3.83 3.82 3.80 3.80 3.79 3.78 3.78 3.77 3.76 3.76 0.10 4
5.86 5.80 5.77 5.75 5.72 5.70 5.69 5.66 5.66 5.65 5.64 5.63 0.05
14.2 14.0 13.9 13.8 13.7 13.7 13.7 13.6 13.6 13.5 13.5 13.5 0.01

1.89 1.88 1.88 1.88 1.88 1.88 1.87 1.87 1.87 1.87 1.87 1.87 0.25
3.24 3.21 3.19 3.17 3.16 3.15 3.14 3.13 3.12 3.12 3.11 3.10 0.10 5
4.62 4.56 4.53 4.50 4.46 4.44 4.43 4.41 4.40 4.39 4.37 4.36 0.05
9.72 9.55 9.47 9.38 9.29 9.24 9.20 9.13 9.11 9.08 9.04 9.02 0.01

1.76 1.76 1.75 1.75 1.75 1.75 1.74 1.74 1.74 1.74 1.74 1.74 0.25
2.87 2.84 2.82 2.80 2.78 2.77 2.76 2.75 2.74 2.73 2.73 2.72 0.10 6
3.94 3.87 3.84 3.81 3.77 3.75 3.74 3.71 3.70 3.69 3.68 3.67 0.05
7.56 7.40 7.31 7.23 7.14 7.09 7.06 6.99 6.97 6.93 6.90 6.88 0.01

1.68 1.67 1.67 1.66 1.66 1.66 1.65 1.65 1.65 1.65 1.65 1.65 0.25
2.63 2.59 2.58 2.56 2.54 2.52 2.51 2.50 2.49 2.48 2.48 2.47 0.01 7
3.51 3.44 3.41 3.38 3.34 3.32 3.30 3.27 3.27 3.25 3.24 3.23 0.05
6.31 6.16 6.07 5.99 5.91 5.86 5.82 5.75 5.74 5.70 5.67 5.65 0.01

1.62 1.61 1.60 1.60 1.59 1.59 1.59 1.58 1.58 1.58 1.58 1.58 0.25
2.46 2.42 2.40 2.38 2.36 2.35 2.34 2.32 2.32 2.31 2.30 2.16 0.10 8
3.22 3.15 3.12 3.08 3.04 2.02 3.01 2.97 2.97 2.95 2.94 2.93 0.05
5.52 5.36 5.28 5.20 5.12 5.07 5.03 4.96 4.95 4.91 4.88 4.86 0.01

1.57 1.56 1.56 1.55 1.55 1.54 1.54 1.53 1.53 1.53 1.53 1.53 0.25
2.34 2.30 2.28 2.25 2.23 2.22 2.21 2.19 2.18 2.17 2.17 2.16 0.10 9
3.01 2.94 2.90 2.86 2.83 2.80 2.79 2.76 2.75 2.73 2.72 2.71 0.05
4.96 4.81 4.73 4.65 4.57 4.52 4.48 4.42 4.40 4.36 4.33 4.31 0.01
Source: As Table B.2: table 18
468 Econometrics far developing countries
Table B.3 Upper percentage points of the F distribution

dffor
denom- df far numerator N 1
inator
N,
Pr 1 2 3 4 5 6 7 8 9 10 11 12

0.25 1.49 1.60 1.60 1.59 1.59 1.58 1.57 1.56 1.56 1.55 1.55 1.54
10 0.10 3.29 2.92 2.73 2.61 2.52 2.46 2.41 2.38 2.35 2.32 2.30 2.28
0.05 4.96 4.10 3.71 3.48 3.33 3.22 3.14 3.07 3.02 2.98 2.94 2.91
0.01 10.0 7.56 6.55 5.99 5.64 5.39 5.20 5.06 4.94 4.85 4.77 4.71

0.25 1.47 1.58 1.58 1.57 1.56 1.55 1.54 1.53 1.53 1.52 1.52 1.51
11 0.10 3.23 2.86 2.66 2.54 2.45 2.39 2.34 2.30 2.27 2.25 2.23 2.21
0.05 4.84 3.98 3.59 3.36 3.20 3.09 3.01 2.95 2.90 2.85 2.82 2.79
0.01 9.65 7.21 6.22 5.67 5.32 5.07 4.89 4.74 4.63 4.54 4.46 4.40

0.25 1.46 1.56 1.56 1.55 1.54 1.53 1.52 1.51 1.51 1.50 1.50 1.49
12 0.10 3.18 2.81 2.61 2.48 2.39 2.33 2.28 2.24 2.21 2.19 2.17 2.15
0.05 4.75 3.89 3.49 3.26 3.11 3.00 2.91 2.85 2.80 2.75 2.72 2.69
0.01 9.33 6.93 5.95 5.41 5.06 4.82 4.64 4.50 4.39 4.30 4.22 4.16

0.25 1.45 1.55 1.55 1.53 1.52 1.51 1.50 1.49 1.49 1.48 1.47 1.47
13 0.10 3.14 2.76 2.56 2.43 2.35 2.28 2.23 2.20 2.16 2.14 2.12 2.10
0.05 4.67 3.81 3.41 3.18 3.03 2.92 2.83 2.77 2.71 2.67 2.63 2.60
0.01 9.07 6.70 5.74 5.21 4.86 4.62 4.44 4.30 4.19 4.10 4.02 3.96

0.25 1.44 1.53 1.53 1.52 1.51 1.50 1.49 1.48 1.47 1.46 1.46 1.45
14 0.10 3.10 2.73 2.52 2.39 2.31 2.24 2.19 2.15 2.12 2.10 2.08 2.05
0.05 4.60 3.74 3.34 3.11 2.96 2.85 2.76 2.70 2.65 2.60 2.57 2.53
0.01 8.86 6.51 5.56 5.04 4.69 4.46 4.28 4.14 4.03 3.94 3.86 3.80

0.25 1.43 1.52 1.52 1.51 1.49 1.48 1.47 1.46 1.46 1.45 1.44 1.44
15 0.10 3.07 2.70 2.49 2.36 2.27 2.21 2.16 2.12 2.09 2.06 2.04 2.02
0.05 4.54 3.68 3.29 3.06 2.90 2.79 2.71 2.64 2.59 2.54 2.51 2.48
0.01 8.86 6.36 5.42 4.89 4.56 4.32 4.14 4.00 3.89 3.80 3.73 3.67

0.25 1.42 1.51 1.51 1.50 1.48 1.47 1.46 1.45 1.44 1.44 1.44 1.43
16 0.10 3.05 2.67 2.46 2.33 2.24 2.18 2.13 2.09 2.06 2.03 2.01 1.99
0.05 4.49 3.63 3.24 3.01 2.85 2.74 2.66 2.59 2.54 2.49 2.46 2.42
0.01 8.53 6.23 5.29 4.77 4.44 4.20 4.03 3.89 3.78 3.69 3.62 3.55

0.25 1.42 1.51 1.50 1.49 1.47 1.46 1.45 1.44 1.43 1.43 1.42 1.41
17 0.10 3.03 2.64 2.44 2.31 2.22 2.15 2.10 2.06 2.03 2.00 1.98 1.96
0.05 4.45 3.59 3.20 2.96 2.81 2.70 2.61 2.55 2.49 2.45 2.41 2.38
0.01 8.40 6.11 5.18 4.67 4.34 4.10 3.93 3.79 3.68 3.59 3.52 3.46

0.25 1.41 1.50 1.49 1.48 1.46 1.45 1.44 1.43 1.42 1.42 1.41 1.40
18 0.10 3.01 2.62 2.42 2.29 2.20 2.13 2.08 2.04 2.00 1.98 1.96 1.93
0.05 4.41 3.55 3.16 2.93 2.77 2.66 2.58 2.51 2.46 2.41 2.37 2.34
0.01 8.29 6.01 5.09 4.58 4.25 4.01 3.84 3.71 3.60 3.51 3.43 3.37

0.25 1.41 1.49 1.49 1.47 1.46 1.44 1.43 1.42 1.41 1.41 1.40 1.40
19 0.10. 2.99 2.61 2.40 2.27 2.18 2.11 2.06 2.02 1.98 1.96 1.94 1.91
0.05 4.38 3.52 3.13 2.90 2.74 2.63 2.54 2.48 2.42 2.38 2.34 2.31
0.01 8.18 5.93 5.01 4.50 4.17 3.94 3.77 3.63 3.52 3.43 3.36 3.30

0.25 1.40 1.49 1.48 1.46 1.45 1.44 1.43 1.42 1.41 1.40 1.39 1.39
20 0.10 2.97 2.59 2.38 2.25 2.16 2.09 2.04 2.00 1.96 1.94 1.92 1.89
0.05 4.35 3.49 3.10 2.87 2.71 2.60 2.51 2.45 2.39 2.35 2.31 2.28
0.01 8.10 5.85 4.94 4.43 4.10 3.87 3.70 3.56 3.46 3.37 3.29 3.23
Estimating simultaneous equation models 469
Table B.3 - Continued

dffor
df far denorninator N 1 denom-
inator
N,
15 20 24 30 40 50 60 100 120 200 500 00 Pr

1.53 1.52 1.52 1.51 1.51 1.50 1.50 1.49 1.49 1.49 1.48 1.48 0.25
2.24 2.20 2.18 2.16 2.13 2.12 2.11 2.09 2.08 2.07 2.06 2.06 0.10 10
2.85 2.77 2.74 2.70 2.66 2.64 2.62 2.59 2.58 2.56 2.55 2.54 0.05
4.56 4.41 4.33 4.25 4.17 4.12 4.08 4.01 4.00 3.96 3.93 3.91 0.01

1.50 1.49 1.49 1.48 1.47 1.47 1.47 1.46 1.46 1.46 1.45 1.45 0.25
2.17 2.12 2.10 2.08 2.05 2.04 2.03 2.00 2.00 1.99 1.98 1.97 0.10 11
2.72 2.65 2.61 2.57 2.53 2.51 2.49 2.46 2.45 2.43 2.42 2.40 0.05
4.25 4.10 4.02 3.94 3.86 3.81 3.78 3.71 3.69 3.66 3.62 3.60 0.01

1.48 1.47 1.46 1.45 1.45 1.44 1.44 1.43 1.43 1.43 1.42 1.42 0.25
2.10 2.06 2.04 2.01 1.99 1.97 1.96 1.94 1.93 1.92 1.91 1.90 0.10 12
2.62 2.54 2.51 2.47 2.43 2.40 2.38 2.35 2.34 2.32 2.31 2.30 0.05
4.01 3.86 3.78 3.70 3.62 3.57 3.54 3.47 3.45 3.41 3.38 3.36 0.01

1.46 1.45 1.44 1.43 1.42 1.42 1.42 1.41 1.41 1.40 1.40 1.40 0.25
2.05 2.01 1.98 1.96 1.93 1.92 1.90 1.88 1.88 1.86 1.85 1.85 0.10 13
2.53 2.46 2.42 2.38 2.34 2.31 2.30 2.26 2.25 2.23 2.22 2.21 0.05
3.82 3.66 3.59 3.51 3.43 3.38 3.34 3.27 3.25 3.22 3.19 3.17 0.01

1.44 1.43 1.42 1.41 1.41 1.40 1.40 1.39 1.39 1.39 1.38 1.38 0.25
2.01 1.96 1.94 1.91 1.89 1.87 1.86 1.83 1.83 1.82 1.80 1.80 0.10 14
2.46 2.39 2.35 2.31 2.27 2.24 2.22 2.19 2.18 2.16 2.14 2.13 0.05
3.66 3.51 3.43 3.35 3.27 3.22 3.18 3.11 3.09 3.06 3.03 3.00 0.01

1.43 1.41 1.41 1.40 1.39 1.39 1.38 1.38 1.37 1.37 1.36 1.36 0.25
1.97 1.92 1.90 1.87 1.85 1.83 1.82 1.79 1.79 1.77 1.76 1.76 0.10 15
2.40 2.33 2.29 2.25 2.20 2.18 2.16 2.12 2.11 2.10 2.08 2.07 0.05
3.52 3.37 3.29 3.21 3.13 3.08 3.05 2.98 2.96 2.92 2.89 2.87 0.01

1.41 1.40 1.39 1.38 1.37 1.37 1.36 1.36 1.35 1.35 1.34 1.34 0.25
1.94 1.89 1.87 1.84 1.81 1.79 1.78 1.76 1.75 1.74 1.73 1.72 0.10 16
2.35 2.28 2.24 2.19 2.15 2.12 2.11 2.07 2.06 2.04 2.02 2.01 0.05
3.41 3.26 3.18 3.10 3.02 2.97 2.93 2.86 2.84 2.81 2.78 2.75 0.01

1.40 1.39 1.38 1.37 1.36 1.35 1.35 1.34 1.34 1.34 1.33 1.33 0.25
1.91 1.86 1.84 1.81 1.78 1.76 1.75 1.73 1.72 1.71 1.69 1.69 0.10 17
2.31 2.23 2.19 2.15 2.10 2.08 2.06 2.02 2.01 1.99 1.97 1.96 0.05
3.31 3.16 3.08 3.00 2.92 2.87 2.83 2.76 2.75 2.71 2.68 2.65 0.01

1.39 1.38 1.37 1.36 1.35 1.34 1.34 1.33 1.33 1.32 1.32 1.32 0.25
1.89 1.84 1.81 1.78 1.75 1.74 1.72 1.70 1.69 1.68 1.67 1.66 0.10 18
2.27 2.19 2.15 2.11 2.06 2.04 2.02 1.98 1.97 1.95 1.93 1.92 0.05
3.23 3.08 3.00 2.92 2.84 2.78 2.75 2.68 2.66 2.62 2.59 2.57 0.01

1.38 1.37 1.36 1.35 1.34 1.33 1.33 1.32 1.32 1.31 1.31 1.30 0.25
1.86 1.81 1.79 1.76 1.73 1.71 1.70 1.67 1.67 1.65 1.64 1.63 0.10 19
2.23 2.16 2.11 2.07 2.03 2.00 1.98 1.94 1.93 1.91 1.89 1.88 0.05
3.15 3.00 2.92 2.84 2.76 2.71 2.67 2.60 2.58 2.55 2.51 2.49 0.01

1.37 1.36 1.35 1.34 1.33 1.33 1.32 1.31 1.31 1.30 1.30 1.29 0.25
1.84 1.79 1.77 1.74 1.71 1.69 1.68 1.65 1.64 1.63 1.62 1.61 0.10 20
2.20 2.12 2.08 2.04 1.99 1.97 1.95 1.91 1.90 1.88 1.86 1.84 0.05
3.09 2.94 2.86 2.78 2.69 2.64 2.61 2.54 2.52 2.48 2.44 2.42 0.01
-·----------------------------
470 Econometrics far developing countries
Table B.3 - Continued

dffor
denom- df for numerator N 1
inator
N,
Pr 1 2 3 4 5 6 7 8 9 10 11 12

0.25 1.40 1.48 1.47 1.45 1.44 1.42 1.41 1.40 1.39 1.39 1.38 1.37
22 0.10 2.95 2.56 2.35 2.22 2.13 2.06 2.01 1.97 1.93 1.90 1.88 1.86
0.05 4.30 3.44 3.05 2.82 2.66 2.55 2.46 2.40 2.34 2.30 2.26 2.23
0.01 7.95 5.72 4.82 4.31 3.99 3.76 3.59 3.45 3.35 3.26 3.18 3.12

0.25 1.39 1.47 1.46 1.44 1.43 1.41 1.40 1.39 1.38 1.38 1.37 1.36
24 0.10 2.93 2.54 2.33 2.19 2.10 2.04 1.98 1.94 1.91 1.88 1.85 1.83
0.05 4.26 3.40 3.01 2.78 2.62 2.51 2.42 2.36 2.30 2.25 2.21 2.18
0.01 7.82 5.61 4.72 4.22 3.90 3.67 3.50 3.36 3.26 3.17 3.09 3.03

0.25 1.38 1.46 1.45 1.44 1.42 1.41 1.39 1.38 1.37 1.37 1.36 1.35
26 0.10 2.91 2.52 2.31 2.17 2.08 2.01 1.96 1.92 1.88 1.86 1.84 1.81
0.05 4.23 3.37 2.98 2.74 2.59 2.47 2.39 2.32 2.27 2.22 2.18 2.15
0.01 7.72 5.53 4.64 4.14 3.82 3.59 3.42 3.29 3.18 3.09 3.02 2.96

0.25 1.38 1.46 1.45 1.43 1.41 1.40 1.39 1.38 1.37 1.36 1.35 1.34
28 0.10 2.89 2.50 2.29 2.16 2.06 2.00 1.94 1.90 1.87 1.84 1.81 1.79
0.05 4.20 3.34 2.95 2.71 2.56 2.45 2.36 2.29 2.24 2.19 2.15 2.12
0.01 7.64 5.45 4.57 4.07 3.75 3.53 3.36 3.23 3.12 3.03 2.96 2.90

0.25 1.38 1.45 1.44 1.42 1.41 1.39 1.38 1.37 1.36 1.35 1.35 1.34
30 0.10 2.88 2.49 2.28 2.14 2.05 1.98 1.93 1.88 1.85 1.82 1.79 1.77
0.05 4.17 3.32 2.92 2.69 2.53 2.42 2.33 2.27 2.21 2.16 2.13 2.09
0.01 7.56 5.39 4.51 4.02 3.70 3.47 3.30 3.17 3.07 2.98 2.91 2.84

0.25 1.36 1.44 1.42 1.40 1.39 1.37 1.36 1.35 1.34 1.33 1.32 1.31
40 0.10 2.84 2.44 2.23 2.09 2.00 1.93 1.87 1.83 1.79 1.76 1.73 1.71
0.05 4.08 3.23 2.84 2.61 2.45 2.34 2.25 2.18 2.12 2.08 2.04 2.00
0.01 7.31 5.18 4.31 3.83 3.51 3.29 3.12 2.99 2.89 2.80 2.73 2.66

0.25 1.35 1.42 1.41 1.38 1.37 1.35 1.33 1.32 1.31 1.30 1.29 1.29
60 0.10 2.79 2.39 2.18 2.04 1.95 1.87 1.82 1.77 1.74 1.71 1.68 1.66
0.05 4.00 3.15 2.76 2.53 2.37 2.25 2.17 2.10 2.04 1.99 1.95 1.92
0.01 7.08 4.98 4.13 3.65 3.34 3.12 2.95 2.82 2.72 2.63 2.56 2.50

0.25 1.34 1.40 1.39 1.37 1.35 1.33 1.31 1.30 1.29 1.28 1.27 1.26
120 0.10 2.75 2.35 2.13 1.99 1.90 1.82 1.77 1.72 1.68 1.65 1.62 1.60
0.05 3.92 3.07 2.68 2.45 2.29 2.17 2.09 2.02 1.96 1.91 1.87 1.83
0.01 6.85 4.79 3.95 3.48 3.17 2.96 2.79 2.66 2.56 2.47 2.40 2.34

0.25 1.33 1.39 1.38 1.36 1.34 1.32 1.31 1.29 1.28 1.27 1.26 1.25
200 0.10 2.73 2.33 2.11 1.97 1.88 1.80 1.75 1.70 1.66 1.63 1.60 1.57
0.05 3.89 3.04 2.65 2.42 2.26 2.14 2.06 1.98 1.93 1.88 1.84 1.80
0.01 6.76 4.71 3.88 3.41 3.11 2.89 2.73 2.60 2.50 2.41 2.34 2.27

0.25 1.32 1.39 1.37 1.35 1.33 1.31 1.29 1.28 1.27 1.25 1.24 1.24
00 0.10 2.71 2.30 2.08 1.94 1.85 1.77 1.72 1.67 1.63 1.60 1.57 1.55
0.05 3.84 3.00 2.60 2.37 2.21 2.10 2.01 1.94 1.88 1.83 1.79 1.75
0.01 6.63 4.61 3.78 3.32 3.02 2.80 2.64 2.51 2.41 2.32 2.25 2.18
Estimating simultaneous equation models 471
Table B.3 - Continued

df far
df for denominator N 1 denom-
inator
N,
15 20 24 30 40 50 60 100 120 200 500 00 Pr

1.36 1.34 1.33 1.32 1.31 1.31 1.30 1.30 1.30 1.29 1.29 1.28 0.25
1.81 1.76 1.73 1.70 1.67 1.65 1.64 1.61 1.60 1.59 1.58 1.57 0.10 22
2.15 2.07 2.03 1.98 1.94 1.91 1.89 1.85 1.84 1.82 1.80 1.78 0.05
2.98 2.83 2.75 2.67 2.58 2.53 2.50 2.42 2.40 2.36 2.33 2.31 0.01

1.35 1.33 1.32 1.31 1.30 1.29 1.29 1.28 1.28 1.27 1.27 1.26 0.25
1.78 1.73 1.70 1.67 1.64 1.62 1.61 1.58 1.57 1.56 1.54 1.53 0.10 24
2.11 2.03 1.98 1.94 1.89 1.86 1.84 1.80 1.79 1.77 1.75 1.73 0.05
2.89 2.74 2.66 2.58 2.49 2.44 2.40 2.33 2.31 2.27 2.24 2.21 0.01

1.34 1.32 1.31 1.30 1.29 1.28 1.28 1.26 1.26 1.26 1.25 1.25 0.25
1.76 1.71 1.68 1.65 1.61 1.59 1.58 1.55 1.54 1.53 1.51 1.50 0.10 26
2.07 1.99 1.95 1.90 1.85 1.82 1.80 1.76 1.75 1.73 1.71 1.69 0.05
2.81 2.66 2.58 2.50 2.42 2.36 2.33 2.25 2.23 2.19 2.16 2.13 0.01

1.33 1.31 1.30 1.29 1.28 1.27 1.27 1.26 1.25 1.25 1.24 1.24 0.25
1.74 1.69 1.66 1.63 1.59 1.57 1.56 1.53 1.52 1.50 1.49 1.48 0.10 28
2.04 1.96 1.91 1.87 1.82 1.79 1.77 1.73 1.71 1.69 1.67 1.65 0.05
2.75 2.60 2.52 2.44 2.35 2.30 2.26 2.19 2.17 2.13 2.09 2.06 0.01

1.32 1.30 1.29 1.28 1.27 1.26 1.26 1.25 1.24 1.24 1.23 1.23 0.25
1.72 1.67 1.64 1.61 1.57 1.55 1.54 1.51 1.50 1.48 1.47 1.46 0.10 30
2.01 1.93 1.89 1.84 1.79 1.76 1.74 1.70 1.68 1.66 1.64 1.62 0.05
2.70 2.55 2.47 2.39 2.30 2.25 2.21 2.13 2.11 2.07 2.03 2.01 0.01

1.30 1.28 1.26 1.25 1.24 1.23 1.22 1.21 1.21 1.20 1.19 1.19 0.25
1.66 1.61 1.57 1.54 1.51 1.48 1.47 1.43 1.42 1.41 1.39 1.38 0.10 40
1.92 1.84 1.79 1.74 1.69 1.66 1.64 1.59 1.58 1.55 1.53 1.51 0.05
2.52 2.37 2.29 2.20 2.11 2.06 2.02 1.94 1.92 1.87 1.83 1.80 0.01

1.27 1.25 1.24 1.22 1.21 1.20 1.19 1.17 1.17 1.16 1.15 1.15 0.25
1.60 1.54 1.51 1.48 1.44 1.41 1.40 1.36 1.35 1.33 1.31 1.29 0.10 60
1.84 1.75 1.70 1.65 1.59 1.56 1.53 1.48 1.47 1.44 1.41 1.39 0.05
2.35 2.20 2.12 2.03 1.94 1.88 1.84 1.75 1.73 1.68 1.63 1.60 0.01

1.24 1.22 1.21 1.19 1.18 1.17 1.16 1.14 1.13 1.12 1.11 1.10 0.25
1.55 1.48 1.45 1.41 1.37 1.34 1.32 1.27 1.26 1.24 1.21 1.19 0.10 120
1.75 1.66 1.61 1.55 1.50 1.46 1.43 1.37 1.35 1.32 1.28 1.25 0.05
2.19 2.03 1.95 1.86 1.76 1.70 1.66 1.56 1.53 1.48 1.42 1.38 0.01

1.23 1.21 1.20 1.18 1.16 1.14 1.12 1.11 1.10 1.09 1.08 1.06 0.25
1.52 1.46 1.42 1.38 1.34 1.31 1.28 1.24 1.22 1.20 1.17 1.14 0.10 200
1.72 1.62 1.57 1.52 1.46 1.41 1.39 1.32 1.29 1.26 1.22 1.19 0.05
2.13 1.97 1.89 1.79 1.69 1.63 1.58 1.48 1.44 1.39 1.33 1.28 0.01

1.22 1.19 1.18 1.16 1.14 1.13 1.12 1.09 1.08 1.07 1.04 1.00 0.25
1.49 1.42 1.38 1.34 1.30 1.26 1.24 1.18 1.17 1.13 1.08 1.00 0.10 00

1.67 1.57 1.52 1.46 1.39 1.35 1.32 1.24 1.22 1.17 1.11 1.00 0.05
2.04 1.88 1.79 1.70 1.59 1.52 1.47 1.36 1.32 1.25 1.15 1.00 0.01

Source: As Table B.2: table 18


472 Econometrics far developing countries
Table B.4 Upper percentage points of the x2 distribution
9SOfo arca
Examp/e
Pr (x2 > 10.85) = 0.95
Pr (x2 > 23.83) = 0.25 for df = 20
Pr (x2 > 31.41) = 0.05

o 10.85 23.83 31.41


Degree Pr
of
free dom 0.995 0.990 0.975 0.950 0.900

1 392704· 10-10 157088·10-9 982069· 10-9 393214·10-8 0.157908


2 0.0100251 0.0201007 0.0506356 0.102587 0.210720
3 0.0717212 0.114832 0.2157950 0.351846 0.584375
4 0.206990 0.297110 0.484419 0.710721 1.063623

5 0.411740 0.554300 0.831211 1.145476 1.61031


6 0.675727 0.872085 1.237347 1.63539 2.20413
7 0.989265 1.239043 1.68987 2.16735 2.83311
8 1.344419 1.646482 2.17973 2.73264 3.48954
9 1.734962 2.087912 2.70039 3.32511 4.16816

10 2.15585 2.55821 3.24697 3.94030 4.86518


11 2.60321 3.05347 3.81575 4.57481 5.57779
12 3.07382 3.57056 4.40379 5.22603 6.30380
13 3.56503 4.10691 5.00874 5.89186 7.04150
14 4.07468 4.66043 5.62872 6.57063 7.78953

15 4.60094 5.22935 6.26214 7.26094 8.54675


16 5.14224 5.81221 6.90766 7.96164 9.31223
17 5.69724 6.40776 7.56418 8.67176 10.08520
18 6.26481 7.01491 8.23075 9.39046 10.86490
19 6.84398 7.63273 8.90655 10.1170 11.6509

20 7.43386 8.26040 9.59083 10.8080 12.4426


21 8.03366 8.89720 10.28293 11.5913 13.2396
22 8.64272 9.54249 10.98230 12.3380 14.0415
23 9.26042 10.19567 11.68850 13.0950 14.8479
24 9.88623 10.8564 12.40110 13.8484 15.6587

25 10.5197 11.5240 13.1197 14.6114 16.4734


26 11.1603 12.1981 13.8439 15.3791 17.2919
27 11.8076 12.8786 14.5733 16.1513 18.1138
28 12.4613 13.5648 15.3079 16.9279 18.9392
29 13.1211 14.2565 16.0471 17.7083 19.7677

30 13.7867 14.9535 16.7908 18.4926 20.5992


40 20.7065 22.1643 24.4331 26.5093 29.0505
50 27.9907 29.7067 32.3574 34.7642 37.6886
60 35.5436 37.4848 40.4817 43.1879 46.4589

70 43.2752 45.4418 48.7576 51.7393 55.3290


80 51.1720 53.5400 57.1532 60.3915 64.2778
90 59.1963 61.7541 65.6466 69.1260 73.2912
lOOa 67.3276 70.0648 74.2219 77.9295 82.3581
Appendix B: Statistical tables 473
Table B.4 - Continued

Degree Pr
of
freedom 0.750 0.500 0.250 0.100 0.050 0.025 0.010 0.005

1 0.1015308 0.454937 1.32330 2.70554 3.84146 5.02389 6.63490 7.87944


2 0.575364 1.38629 2.77259 4.60517 5.99147 7.37776 9.21034 10.5966
3 1.212534 2.36597 4.10835 6.25139 7.81473 9.34840 11.3449 12.8381
4 1.92255 3.35670 5.38527 7.77944 9.48773 11.1433 13.2767 14.8602

5 2.67460 4.35146 6.62568 9.23635 11.0705 12.8325 15.0863 16.7496


6 3.45460 5.34812 7.84080 10.6446 12.5916 14.4494 16.8119 18.5476
7 4.25485 6.34581 9.03715 12.0170 14.0671 16.0128 18.4753 20.2777
8 5.07064 7.34412 10.2188 13.3616 15.5073 17.5346 20.0902 21.9550
9 5.89883 8.34283 11.3887 14.6837 16.9190 19.0228 21.6660 23.5893

10 6.73720 9.34182 12.5489 15.9871 18.3070 20.4831 23.2093 25.1882


11 7.58412 10.3410 13.7007 17.2750 19.6751 21.9200 24.7250 26.7569
12 8.43842 11.3403 14.8454 18.5494 21.0261 23.3367 26.2170 28.2995
13 9.29906 12.3398 15.9839 19.8119 22.3621 24.7356 27.6883 29.8194
14 10.1653 13.3393 17.1170 21.0642 23.6848 26.1190 29.1413 31.3193

15 11.0365 14.3389 18.2451 22.3072 24.9958 27.4884 30.5779 32.8031


16 11.9122 15.3385 19.3688 23.5418 26.2962 28.8454 31.9999 34.2672
17 12.7919 16.3381 20.4887 24.7690 27.5871 30.1910 33.4087 35.7185
18 13.6753 17.3379 21.6049 25.9894 28.8693 31.5264 34.8053 37.1564
19 14.5620 18.3376 22.7178 27.2036 30.1435 32.8523 36.1908 38.5822

20 15.4518 19.3374 23.8277 28.4120 31.4104 34.1696 37.5662 39.9968


21 16.3444 20.3372 24.9348 29.6151 32.6705 35.4789 38.93210 41.4010
22 17.2396 21.3370 26.0393 30.8133 33.9244 36.7807 40.28940 42.7956
23 18.1373 22.3369 27.1413 32.0069 35.1725 38.0757 41.63840 44.1813
24 19.0372 23.3367 28.2412 33.1963 36.4151 39.3641 42.97980 45.5585

25 19.9393 24.3366 29.3389 34.3816 37.6525 40.6465 44.3141 46.9278


26 20.8434 25.3364 30.4345 35.5631 38.8852 41.9232 45.6417 48.2899
27 21.7494 26.3363 31.5284 36.7412 40.1133 43.1944 46.9630 49.6449
28 22.6572 27.3363 32.6205 37.9159 41.3372 44.4607 48.2782 50.9933
29 23.5666 28.3362 33.7109 39.0875 42.5569 45.7222 49.5879 52.3356

30 24.4776 29.336 34.7998 40.256 43.7729 46.9792 50.8922 53.6720


40 33.6603 39.3354 45.6160 51.8050 55.7585 59.3417 63.6907 66.7659
50 42.9421 49.3349 56.3336 63.1671 67.5048 71.4202 76.1539 79.4900
60 52.2938 59.3347 66.9814 74.3970 79.0819 83.2976 88.3794 91.9517

70 61.6983 69.3344 77.5766 85.5271 90.5312 95.0231 100.425 104.215


80 71.1445 79.3343 88.1303 96.5782 101.879 106.629 112.329 116.321
90 80.6247 89.3342 98.6499 107.565 113.145 118.136 124.116 128.299
lOOa 90.1332 99.3341 109.141 118.498 124.342 129.561 135.807 140.169

So urce: Abridged from same source as Table B.2: table 8

Notes: For df greater than 100 the expansion: [sqr]2x2 - [sqr](2k - 1) = Z follows the standardised normal distribution.
where k represents the degrees of freedom.
474 Econometrics far developing countries
Table B.5 Durbin-Watson dstatistic: significance points of dL and du at 0.05 level of signiticance
Example
If n = 40 and k' = 4, d¿ = l.28S and du = 1.721. If a computed d value is less !han 1.28S, there is evidence of positive first-or<
serial correlation, if it is greater than 1.721 there is no evidence of positive first-order serial correlation, but if d lies between 1
lower and the upper limit, there is conclusive evidence regarding the presence ar absence of positive first-order serial correlation

k' = 1 k' = 2 k' = 3 k' = 4 k' = 5

n dL du dL du dL du dL du dL du

6 0.610 1.400
7 0.700 l.3S6 0.467 1.896
8 0.763 1.332 O.SS9 1.777 0.368 2.287
9 0.824 1.320 0.629 1.699 0.4SS 2.128 0.296 2.S88
10 0.879 1.320 0.697 1.641 O.S2S 2.016 0.376 2.414 0.243 2.822
11 0.927 1.324 0.6S8 1.604 0.59S 1.928 0.444 2.283 0.316 2.64S
12 0.971 1.331 0.812 1.579 0.6S8 1.864 0.Sl2 2.177 0.379 2.S06
13 1.010 1.340 0.861 1.562 0.71S 1.816 0.S74 2.094 0.44S 2.390
14 l.04S l.3SO 0.90S l.SSl 0.767 1.779 0.632 2.030 o.sos 2.296
lS 1.077 1.361 0.946 1.543 0.814 l.?SO 0.68S 1.977 O.S62 2.220
16 1.106 1.371 0.982 1.539 0.87S 1.728 0.734 l.93S 0.6lS 2.1S7
17 1.133 1.381 l.OlS 1.536 0.897 1.710 0.779 1.900 0.664 2.104
18 l.lS8 1.391 1.046 l.S3S 0.933 1.696 0.820 1.872 0.710 2.060
19 1.180 1.401 1.074 1.536 0.967 l.68S 0.8S9 1.848 0.7S2 2.023
20 1.201 1.411 1.100 1.537 0.998 1.676 0.894 1.828 0.792 1.991
21 1.221 1.420 l.l2S 1.538 1.026 1.669 0.927 1.812 0.829 1.964
22 1.239 1.429 1.147 1.S41 I.OS3 1.664 0.9S8 1.797 0.863 1.940
23 l.2S7 1.437 1.168 l.S43 1.078 1.660 0.986 1.78S 0.89S 1.920
24 1.273 1.446 1.188 1.546 1.101 l.6S6 1.013 1.77S 0.92S 1.902
2S 1.288 l.4S4 1.206 l.SSO 1.123 l.6S4 1.038 1.767 0.9S3 1.886
26 1.302 1.461 1.224 l.SS3 1.143 l.6S2 1.062 1.7S9 0.979 1.873
27 1.316 1.469 1.240 l.SS6 1.162 l.6Sl 1.084 1.7S3 1.004 1.861
28 1.328 1.476 I.2SS l.S60 1.181 l.6SO 1.104 1.747 1.028 l.8SO
29 1.341 1.483 1.270 l.S63 1.198 l.6SO 1.124 1.743 l.OSO 1.841
30 l.3Sl 1.489 1.284 1.S67 1.214 l.6SO 1.143 1.739 1.071 1.833
31 1.363 1.496 1.297 1.570 1.229 l.6SO 1.160 I.73S 1.090 l.82S
32 1.373 1.502 1.309 1.574 1.244 I.6SO 1.177 1.732 1.109 1.819
33 1.383 l.S08 1.321 l.S77 l.2S8 1.6Sl 1.193 1.730 1.127 1.813
34 1.393 1.514 1.333 1.580 1.271 I.6S2 1.208 1.728 1.144 1.808
3S 1.402 1.S19 1.343 I.S84 1.283 I.6S3 1.222 1.726 1.160 1.803
36 1.411 l.S2S l.3S4 1.S87 I.29S I.6S4 1.236 1.724 l.17S 1.799
37 1.419 I.S30 1.364 I.S90 1.307 I.6SS 1.249 1.723 1.190 l.79S
38 1.427 l.S3S 1.373 1.594 1.318 I.6S6 1.261 1.722 1.204 1.792
39 l.43S l.S40 1.382 1.597 1.328 l.6S8 1.273 1.722 1.218 1.789
40 1.442 1.544 1.391 1.600 1.338 l.6S9 I.28S 1.721 1.230 1.786
4S 1.47S I.S66 1.430 1.61S 1.383 1.666 1.336 1.720 1.287 1.776
so I.S03 l.58S 1.462 1.628 1.421 1.674 1.378 1.721 I.33S 1.771
SS I.S28 1.601 1.490 1.641 l.4S2 1.681 1.414 1.724 1.374 1.768
60 l.S49 1.616 I.Sl4 I.6S2 1.480 1.689 1.444 1.727 1.408 1.767
6S l.S67 1.629 1.536 1.662 1.503 1.696 1.471 1.731 1.438 1.767
70 1.583 1.641 l.SS6 1.672 l.52S 1.703 1.494 1.73S 1.464 1.768
7S 1.598 l.6S2 1.571 1.680 I.S43 1.709 l.SlS 1.739 1.487 1.770
80 1.611 1.662 I.S86 1.688 I.S60 1.71S l.S34 1.743 1.507 1.772
8S 1.624 1.671 1.600 1.696 I.S7S 1.721 l.SSO 1.747 l.S2S 1.774
90 I.63S 1.679 1.612 1.703 I.S89 1.726 l.S66 l.7Sl l.S42 1.776
9S I.64S 1.687 1.623 1.709 1.602 1.732 1.S79 I.7SS l.SS7 1.778
100 I.6S4 1.694 1.634 1.71S 1.613 1.736 l.S92 1.7S8 1.S71 1.780
lSO 1.720 1.746 1.706 1.760 1.693 1.774 1.679 1.788 l.66S 1.802
200 I.7S8 1.778 1.748 1.789 1.738 1.799 1.728 1.810 1.718 1.820
Appendix B: Statistical tables 475
Table B.5 - Continued

k' = 6 k' = 7 k' = 8 k' = 9 k' = 10

n dL du du dL du dL du dL dL du

6
7
8
9
10
11 0.203 3.005
12 0.268 2.835 0.171 3.149
13 0.328 2.692 0.230 2.985 0.147 3.266
14 0.389 2.572 0.286 2.848 0.200 3.111 0.127 3.360
15 0.447 2.472 0.343 2.727 0.251 2.979 0.175 3.216 0.111 3.438
16 0.502 2.388 0.398 2.624 0.304 2.860 0.222 3.090 0.155 3.304
17 0.554 2.318 0.451 2.537 0.356 2.757 0.272 2.975 0.198 3.184
18 0.603 2.257 0.502 2.461 0.407 2.667 0.321 2.873 0.244 3.073
19 0.649 2.206 0.549 2.396 0.456 2.589 0.369 2.783 0.290 2.974
20 0.692 2.162 0.595 2.339 0.502 2.521 0.416 2.704 0.336 2.885
21 0.732 2.124 0.637 2.290 0.547 2.460 0.461 2.633 0.380 2.806
22 0.769 2.090 0.677 2.246 0.588 2.407 0.504 2.571 0.424 2.734
23 0.804 2.061 0.715 2.208 0.628 2.360 0.545 2.514 0.465 2.670
24 0.837 2.035 0.751 2.174 0.666 2.318 0.584 2.464 0.560 2.613
25 0.868 2.012 0.784 2.144 0.702 2.280 0.621 2.419 0.544 2.560
26 0.897 1.992 0.816 2.117 0.735 2.246 0.657 2.379 0.581 2.513
27 0.925 1.974 0.845 2.093 0.767 2.216 0.691 2.342 0.616 2.470
28 0.951 1.958 0.874 2.071 0.798 2.188 0.723 2.309 0.650 2.431
29 0.975 1.944 0.900 2.052 0.826 2.164 0.753 2.278 0.682 2.396
30 0.998 1.931 0.926 2.034 0.854 2.141 0.782 2.251 0.712 2.363
31 1.020 1.920 0.950 2.018 0.879 2.120 0.810 2.226 0.741 2.333
32 1.041 1.909 0.972 2.004 0.904 2.102 0.836 2.203 0.769 2.306
33 1.061 1.900 0.994 1.991 0.927 2.085 0.861 2.181 0.795 2.281
34 1.080 1.891 1.015 1.979 0.950 2.069 0.885 2.162 0.821 2.257
35 1.097 1.884 1.034 1.967 0.971 2.054 0.908 2.144 0.845 2.236
36 1.114 1.877 1.053 1.957 0.991 2.041 0.930 2.127 0.868 2.216
37 1.131 1.870 1.071 1.948 1.011 2.029 0.951 2.112 0.891 2.198
38 1.146 1.864 1.088 1.939 1.029 2.017 0.970 2.098 0.912 2.180
39 1.161 1.859 1.104 1.932 1.047 2.007 1.990 2.085 0.932 2.164
40 1.175 1.854 1.120 1.924 1.064 1.997 1.008 2.072 0.952 2.149
45 1.238 1.835 1.189 1.895 1.139 1.958 1.089 2.022 1.038 2.088
50 1.291 1.822 1.246 1.875 1.201 1.930 1.156 1.986 1.110 2.044
55 1.334 1.814 1.294 1.861 1.253 1.909 1.212 1.959 1.170 2.010
60 1.372 1.808 1.335 1.850 1.298 1.894 1.260 1.939 1.222 1.984
65 1.404 1.805 1.370 1.843 1.336 1.882 1.301 1.923 1.266 1.964
70 1.433 1.802 1.401 1.837 1.369 1.873 1.337 1.910 1.305 1.948
75 1.458 1.801 1.428 1.834 1.399 1.867 1.369 1.901 1.339 1.935
80 1.480 1.801 1.453 1.831 1.425 1.861 1.397 1.893 1.369 1.925
85 1.500 1.801 1.474 1.829 1.448 1.857 1.422 1.886 1.396 1.916
90 1.518 1.801 1.494 1.827 1.469 1.854 1.445 1.881 1.420 1.909
95 1.535 1.802 1.512 1.827 1.489 1.852 1.465 1.877 1.442 1.903
100 1.550 1.803 1.528 1.826 1.506 1.850 1.484 1.874 1.462 1.898
150 1.651 1.817 1.637 1.832 1.622 1.847 1.608 1.862 1.594 1.877
200 1.707 1.831 1.697 1.841 1.686 1.852 1.675 1.863 1.665 1.874
476 Econometrics far developing countries
Table B.5 - Continued

k' = 11 k' = 12 k' = 13 k' =14 k' = 15

n dL du dL du dL du dL du dL du

6
7
8
9
10
11
12
13
14
15
16 0.098 3.503
17 0.138 3.378 0.087 3.557
18 0.177 3.265 0.123 3.441 0.078 3.603
19 0.220 3.159 0.160 3.335 0.111 3.496 0.070 3.642
20 0.263 3.063 0.200 3.234 0.145 3.395 0.100 3.542 0.063 3.676
21 0.307 2.976 0.240 3.141 0.182 3.300 0.132 3.448 0.091 3.583
22 0.349 2.897 0.281 3.057 0.220 3.211 0.166 3.358 0.120 3.495
23 0.391 2.826 0.322 2.979 0.259 2.128 0.202 3.272 0.153 3.409
24 0.431 2.761 0.362 2.908 0.297 2.053 0.239 3.193 0.186 3.327
25 0.470 2.702 0.400 2.844 0.335 2.983 0.275 3.119 0.221 3.251
26 0.508 2.649 0.438 2.784 0.373 2.919 0.312 3.051 0.256 3.179
27 0.544 2.600 0.475 2.730 0.409 2.859 0.348 2.987 0.291 3.112
28 0.578 2.555 0.510 2.680 0.445 2.805 0.383 2.928 0.325 3.050
29 0.612 2.515 0.544 2.634 0.479 2.755 0.418 2.874 0.359 2.992
30 0.643 2.477 0.577 2.592 0.512 2.708 0.451 2.823 0.392 2.937
31 0.674 2.443 0.608 2.553 0.545 2.665 0.484 2.776 0.425 2.887
32 0.703 2.411 0.638 2.517 0.576 2.625 0.515 2.733 0.457 2.840
33 0.731 2.382 0.668 2.484 0.606 2.588 0.546 2.692 0.488 2.796
34 0.758 2.355 0.695 2.454 0.634 2.554 0.575 2.654 0.518 2.754
35 0.783 2.330 0.722 2.425 0.662 2.521 0.604 2.619 0.547 2.716
36 0.808 2.306 0.748 2.398 0.689 2.492 0.631 2.586 0.575 2.680
37 0.831 2.285 0.772 2.374 0.714 2.464 0.657 2.555 0.602 2.646
38 0.854 2.265 0.796 2.351 0.739 2.438 0.683 2.526 0.628 2.614
39 0.875 2.246 0.819 2.329 0.763 2.413 0.707 2.499 0.653 2.585
40 0.896 2.228 0.840 2.309 0.785 2.391 0.731 2.473 0.678 2.557
45 0.988 2.156 0.938 2.225 0.887 2.296 0.838 2.367 0.788 2.439
50 1.064 2.103 1.019 2.163 0.973 2.225 0.927 2.287 0.882 2.350
55 1.129 2.062 1.087 2.116 1.045 2.170 1.003 2.225 0.961 2.281
60 1.184 2.031 1.145 2.079 1.106 2.127 1.068 2.177 1.029 2.227
65 1.231 2.006 1.195 2.049 1.160 2.093 1.124 2.138 1.088 2.183
70 1.272 1.986 1.239 2.026 1.206 2.066 1.172 2.106 1.139 2.148
75 1.308 1.970 1.277 2.006 1.247 2.043 1.215 2.080 1.184 2.118
80 1.340 1.957 1.311 1.991 1.283 2.024 1.253 2.059 1.224 2.093
85 1.369 1.946 1.342 1.977 1.315 2.009 1.287 2.040 1.260 2.073
90 1.395 1.937 1.369 1.966 1.344 1.995 1.318 2.025 1.292 2.055
95 1.418 1.929 1.394 1.956 1.370 1.984 1.345 2.012 1.321 2.040
100 1.439 1.923 1.416 1.948 1.393 1.974 1.371 2.000 1.347 2.026
150 1.579 1.892 1.564 1.908 1.550 1.924 1.535 1.940 1.519 1.956
200 1.654 1.885 1.643 1.896 1.632 1.908 1.621 1.919 1.610 1.931
Appendix B: Statistical tables 477
rabie B.5 - Continued

k' = 16 k' = 17 k' = 18 k' = 19 k' = 20

dL du dL du dL du dL du dL du

6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21 0.058 3.705
22 0.083 3.619 0.052 3.731
23 0.110 3.535 0.076 3.650 0.048 3.7S3
24 0.141 3.454 0.101 3.S72 O.ü70 3.678 0.044 3.773
25 0.172 3.376 0.130 3.494 0.094 3.604 0.06S 3.702 0.041 3.790
26 0.205 3.303 0.160 3.420 0.210 3.S31 0.087 3.632 0.060 3.724
27 0.238 3.233 0.191 3.349 0.149 3.460 0.112 3.S63 0.081 3.6S8
28 0.271 3.168 0.222 3.283 0.178 3.392 0.138 3.49S 0.104 3.S92
29 0.30S 3.107 0.2S4 3.219 0.208 3.327 0.166 3.431 0.129 3.S28
30 0.337 3.0SO 0.286 3.160 0.238 3.266 0.19S 3.368 0.156 3.465
31 0.370 2.996 0.317 3.103 0.269 3.208 0.224 3.309 0.183 3.406
32 0.401 2.946 0.349 3.0SO 0.299 3.1S3 0.2S3 3.252 0.211 3.348
33 0.432 2.899 0.379 3.000 0.329 3.100 0.283 3.198 0.239 3.293
34 0.462 2.854 0.409 2.9S4 0.3S9 3.0Sl 0.312 3.147 0.267 3.240
3S 0.492 2.813 0.439 2.910 0.388 3.005 0.340 3.099 0.295 3.190
36 0.520 2.774 0.467 2.868 0.417 2.961 0.369 3.053 0.323 3.142
37 0.S48 2.738 0.49S 2.829 0.44S 2.920 0.397 3.009 0.351 3.097
38 0.S7S 2.703 O.S22 2.792 0.472 2.880 0.424 2.968 0.378 3.054
39 0.600 2.671 0.549 2.757 0.499 2.843 0.4Sl 2.929 0.404 3.013
40 0.626 2.641 0.57S 2.724 0.S2S 2.808 0.477 2.892 0.430 2.974
4S 0.740 2.512 0.692 2.S86 0.644 2.659 0.S98 2.733 O.SS3 2.807
so 0.836 2.414 0.792 2.479 0.747 2.S44 0.703 2.610 0.660 2.67S
SS 0.919 2.338 0.877 2.396 0.836 2.4S4 0.79S 2.512 0.7S4 2.S71
60 0.990 2.278 0.9Sl 2.330 0.913 2.382 0.874 2.434 0.836 2.487
6S 1.0S2 2.229 1.016 2.276 0.980 2.323 0.944 2.371 0.908 2.419
70 1.lOS 2.189 1.072 2.232 1.038 2.27S 1.005 2.318 0.971 2.362
7S 1.1S3 2.156 1.121 2.19S 1.090 2.23S 1.0S8 2.275 1.027 2.317
80 1.195 2.129 l.l6S 2.165 1.136 2.201 1.106 2.238 1.076 2.275
85 1.232 2.lOS 1.205 2.139 1.177 2.172 1.149 2.206 1.121 2.241
90 1.266 2.08S 1.240 2.116 1.213 2.148 1.187 2.179 1.160 2.211
9S 1.296 2.068 1.271 2.097 1.247 2.126 1.222 2.1S6 1.197 2.186
00 1.324 2.053 1.301 2.080 1.277 2.108 1.2S3 2.13S 1.229 2.164
50 1.S04 1.972 1.489 1.989 1.474 2.006 1.4S8 2.023 1.443 2.040
:oo 1.S99 1.943 1.S88 1.9SS 1.S76 1.967 1.S6S 1.979 1.SS4 1.991

;ource: This table is an extension of the original Durbin-Watson table and is reproduced from N.E. Savín and K.J. White 'The
)urbin-Watson Test for Serial Correlation with Extreme Small Samples ar Many Regressors', Econometrica, 45, November 1977,
989-96 and as corrected by R.W. Farebrother, Econometrica, 48, September 1980, lSS4.

l/otes:

! = number of observations
;' = number of explanatory variables excluding the constant term
478 Econometrics for developing countries
Table B.6 Critica! values of runs in the runs test

Example
In a sequence of 30 observations consisting of 20 + signs (= N 1) and 10 - signs (= N2), the
critica! values of runs at the 0.05 leve! of significance are 9 and 20, as shown by Table B.6(a)
and (b), respectively. Therefore, if in an application it is found that the number of runs is
equal to or less than 9 or equal or greater than 20, one can reject (at the 0.05 leve! of signif-
icance) the hypothesis that the observed sequence is random.
(a)
Nz

N, 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
2 2 2 2 2 2 2 2 2 2
3 2 2 2 2 2 2 2 2 2 3 3 3 3 3
4 2 2 2 3 3 3 3 3 3 3 3 4 4 4 4 4
5 2 2 3 3 3 3 3 4 4 4 4 4 4 4 5 5 5
6 2 2 3 3 3 3 4 4 4 4 5 5 5 5 5 5 6 6
7 2 2 3 3 3 4 4 5 5 5 5 5 6 6 6 6 6 6
8 2 3 3 3 4 4 5 5 5 6 6 6 6 6 7 7 7 7
9 2 3 3 4 4 5 5 5 6 6 6 7 7 7 7 8 8 8
10 2 3 3 4 5 5 5 6 6 7 7 7 7 8 8 8 8 9
11 2 3 4 4 5 5 6 6 7 7 7 8 8 8 9 9 9 9
12 2 2 3 4 4 5 6 6 7 7 7 8 8 8 9 9 9 10 10
13 2 2 3 4 5 5 6 6 7 7 8 8 9 9 9 10 10 10 10
14 2 2 3 4 5 5 6 7 7 8 8 9 9 9 10 10 10 11 11
15 2 3 3 4 5 6 6 7 7 8 8 9 9 10 10 11 11 11 12
16 2 3 4 4 5 6 6 7 8 8 9 9 10 10 11 11 11 12 12
17 2 3 4 4 5 6 7 7 8 9 9 10 10 11 11 11 12 12 13
18 2 3 4 5 5 6 7 8 8 9 9 10 10 11 11 12 12 13 13
19 2 3 4 5 6 6 7 8 8 9 10 10 11 11 12 12 13 13 13
20 2 3 4 5 6 6 7 8 9 9 10 10 11 12 12 13 13 14 14
Estimating simultaneous equation models 479
(b)
N1

N¡ 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
2
3
4 9 9
5 9 10 10 11 11
6 9 10 11 12 12 13 13 13 13
7 11 12 13 13 14 14 14 14 15 15 15
8 11 12 13 14 14 15 15 16 16 16 16 17 17 17 17
9 13 14 14 15 16 16 16 17 17 18 18 18 18 18 18
10 13 14 15 16 16 17 17 18 18 18 19 19 19 20 20
11 13 14 15 16 17 17 18 19 19 19 20 20 20 21 21
12 13 14 16 16 17 18 19 19 20 20 21 21 21 22 22
13 15 16 17 18 19 19 20 20 21 21 22 22 23 23
14 15 16 17 18 19 20 20 21 22 22 23 23 23 24
15 15 16 18 18 19 20 21 22 22 23 23 24 24 25
16 17 18 19 20 21 21 22 23 23 24 25 25 25
17 17 18 19 20 21 22 23 23 24 25 25 26 26
18 17 18 19 20 21 22 23 24 25 25 26 26 27
19 17 18 20 21 22 23 23 24 25 26 26 27 27
20 17 18 20 21 22 23 24 25 25 26 27 27 28
Source: Sidney Siegel, Nonparametric Statistic for the Behavioral Science, New York:
McGraw-Hill, 1956, table F, pp. 252-3. The table has been adapted by Siegel from the
original source: Frieda S. Swed and C. Eisenhart, 'Tables for Testing Randomness of
Grouping in a Sequence of Altematives', Annals of Mathematical Statistics, 14, 1943.
Notes: Table B.6(a) and (b) give the critica! values of runs n for various values of N 1(+
symbol) and Nz(-symbol). For the one sample runs test, any value of n which is equal to
or smaller than that shown in Table B.6(a) or equal to or larger than shown in Table
B.6(b) is significant at the 0.05 leve!.
480 Econometrics far developing countries
Table B. 7 Critical values for Dickey-Fuller test
Sample size Probability of a larger value
10% 5% 1%

<1>1: (13 1,13 3) = (0,1) in x, = 13 1 + 13 3 X,_1 + e,


25 4.12 5.18 7.88
50 3.94 4.86 7.06
100 3.86 4.71 6.70
00 3.78 4.59 6.43
<1>2: Cl31,l32.l33) = 0,0,1) in x, = 131 + 132 t + l33X,_1 + e,
25 4.67 5.68 8.21
50 4.31 5.13 7.02
100 4.16 4.88 6.50
4.03 4.68 6.09
<1>3: Cl31,l32,l33) = 131,0,1) in X1 = 131 + 132 t + l33X,_1 + e,
25 5.91 7.24 10.61
50 5.61 6.73 9.31
100 5.47 6.49 8.73
5.34 6.25 8.27
T 3: 13 3 = 1 in x, = 13 1 + 13 2 t + 133X,_ 1 + e,
25 -1.14 -0.80 -0.15
50 -1.19 -0.87 -0.24
100 -1.22 -0.90 -0.28
00 -1.25 -0.94 -0.33
Source: Fuller (1976) and Dickey and Fuller (1981: 1063)
References

Aldrich, John H. and Nelson, Forrest D. (1984) Linear Probability, Logit, and
Probit Models, Beverly Hills, CA: Sage.
Banerjee, Anindya, Dolado, Juan J., Galbraith, John J. and Hendry, David F.
(1993) Cointegration, Error Correction, and the Econometric Analysis of Non-
stationary Data, Oxford: Oxford University Press.
Barten, A.P. (1985) 'Het voorgeborchte der econometrie: het parelsnoer van
Engel', Tijdschrift voor economie en management 30 (3-4): 453-74.
Box, G.E.P. and Jenkins G.M. (1970) Time Series Analysis, Forecasting and
Control, San Francisco: Holden-Day.
Carr, E.H. ([1961], 1990) What is History?, ed. R.W. Davies, Harmondsworth:
Penguin Books.
Chambers, John M., Cleveland, William S., Kleiner, Beat and Tukey, Paula S.
(1983) Graphical Methods for Data Analysis, Pacific Grove, CA: Wadsworth &
Brooks/Cole Publishing Company Advanced Books & Software.
Charemza, Wojciech W. and Deadman Derek F. (1992) New Directions in Econo-
metric Practice: General to Specific Modelling, Cointegration, Aldershot, Hants.:
Edward Elgar.
Chenery, Hollis B. and Strout, William (1966) 'Foreign Assistance and Economic
Development', American Economic Review 66: 679-733.
Collier, P., Radwan, S. and Wangwe, S. (with Wagner, A.) (1986) Labour and
Poverty in Rural Tanzania: Ujamaa and Rural Development in the United
Republic of Tanzania, Oxford: Clarendon Press.
Dasgupta, P. (1993) An Jnquiry into Well-Being and Destitution, Oxford: Oxford
University Press.
Davies, Richard B. (1994) 'From Cross-sectional to Longitudinal Analysis', in
Analysing Social and Political Change: A Casebook of Methods, ed. Angela Dale
and Richard B. Davies, London: Sage Publications.
Demaris, Alfred (1992) Logit Modelling: Practica! Applications, Beverly Hills, CA:
Sage Publications.
Diaconis, P. (1985) 'Theories of Data Analysis: From Magical Thinking Through
Classical Statistics', in David C. Hoaglin, F. Mosteller and J. Tukey (eds),
Exploring Data Tables, Trends and Shapes, New York: John Wiley.
Dickey, D.A. and Fuller, W.A. (1981) 'Likelihood Ration Statistics for Autore-
gressive Time Series With a Unit Root', Econometrica 49: 12-26.
Emerson, John D. and Hoaglin, David C. (1985) 'Resistant Multiple Regression,
One Variable at a Time' in David C. Hoaglin, Frederick Mosteller and John W.
Tukey (eds), Exploring Data Tables, Trends, and Shapes, New York: John Wiley.
Emerson, John D. and Strenio, Judith (1983) 'Boxplots and Batch Comparison',
in David C. Hoaglin Frederick Mosteller and John W. Tukey (eds ), Under-
standing Robust and Exploratory Data Analysis, New York: John Wiley.
482 Econometrics far developing countries
Engle, Robert F. and Granger, C.W.J. (1987) 'Cointegration and Error Correction:
Representation, Estimation and Testing', Econometrica 55 (2): 251-76.
Engle, Robert F., Hendry, D.F. and Richard, J.F. (1983) 'Exogeneity', Econo-
metrica 55: 277-304.
Friedman, Milton (1970) The Controversy in Monetary Theory: The First Winicott
Memorial Lecture, 16 September, JEA Occasional Paper no. 33. London:
Institute of Economic Affairs.
Fry, M.J. (1988) Money, Interest, and Banking in Economic Development,
Baltimore, MD: Johns Hopkins University Press.
Fuller, W.A. (1976) Introduction to Statistical Time Series, New York: John Wiley.
Gatt, J. (1995) An Econometric Analysis of the Determinants of the Maltese Exports
of Manufacturers, ISS Working Paper no. 198, The Hague: Institute of Social
Studies.
Giere, Ronald N. (1991) Understanding Scientific Reasoning, Fort Worth, IL: Holt,
Rinehart & Winston.
Gilbert, Christopher (1990) 'Professor Hendry's Econometric Methodology', in
C.W.J. Granger (ed.), Modelling Economic Series, Oxford: Clerendon Press.
Goldberger, Arthur S. (1964) Econometric Theory, New York: John Wiley.
Goldberger, Arthur S. (1991) A Course in Econometrics, Cambridge, MA: Harvard
University Press.
Gould, S. J. (1996) Full House: The Spread of Excellence from Plato to Darwin,
New York: Harmony Books.
Granger, C.W.J. (ed.) (1990) Modelling Economic Series, Oxford: Clarendon Press.
Granger, C.W.J. and Newbold, P. (1974) 'Spurious Regressions in Econometrics'
Journal of Econometrics 2: 111-20.
Granger, C.W.J. and Newbold, P. (1977) Forecasting Economic Time Series, New
York: Academic Press; 2nd edition 1986.
Gregory, C.A. and Altman, J.C. (1989) Observing the Economy, London: Routledge.
Griffin, Keith (1970) 'Foreign Capital, Domestic Savings and Economic
Development', Bulletin of the Oxford University Institute of Economics and
Statistics 32: 99-112.
Griffin, Keith (1971) 'Reply', Bulletin of the Oxford University Institute of
Economics and Statistics 33: 156-61.
Griffin, Keith and Enos, John (1971) 'Foreign Assistance: Objectives and
Consequences', Economics of Development and Cultural Change 18: 313-27.
Grilli, E. and Yang, M.C. (1988) 'Primary Commodity Prices, Manufactured Goods
Prices and the Terms of Trade of Developing Countries: What the Long Run
Shows', World Bank Economic Review 2: 1-47.
Gujarati, D. (1988), Basic Econometrics, 2nd edn, New York: McGraw-Hill.
Hamilton, James D. (1994) Time Series Analysis, Princeton, NJ: Princeton
University Press.
Hamilton, Lawrence (1990) Modern Data Analysis: A First Course in Applied
Statistics, Pacific Grove, CA: Brooks Cole.
Hamilton, Lawrence C. (1992) Regression with Graphics: A Second Course in
Applied Statistics, Pacific Grove, CA: Brooks Cole.
Harriss, B. (1990) 'The Intrafamily Distribution of Hunger in South Asia', in Jean
Dreze and Amaryta Sen (eds), The Political Economy of Hunger, vol. 1,
Entitlement and Well-being, Oxford: Clarendon Press.
Heckman, J.J. (1992) 'Haavelmo and the Birth of Modern Econometrics: A
Review of The History of Econometric Ideas by Mary Morgan', Journal of
Econometric Literature 30: 876-86.
Helmers, F.L.C.H. (1988) 'Real Exchange Rate Indexes', in R. Dornsbuch and
F.L.C.H. Helmers (eds), The Open Economy: Tools far Policy Makers in
Developing Countries, Oxford: Oxford University Press.
References 483
Hoaglin, David C. (1983) 'Letter Values: a Set of Selected Order Statistics', in
David C. Hoaglin, F. Mosteller and J. Tukey Understanding Robust and
Exploratory Data Analysis, New York: John Wiley, pp. 33-57.
Hoaglin, David C. (1985) 'Using Quintiles to Study Shapes', in David C. Hoaglin,
F. Mosteller and J. Tukey, Exploring Data Tables, Trends and Shapes, New York:
John Wiley.
Hoaglin, David C., Mosteller, F. and Tukey, J. (1983) Understanding Robust and
Exploratory Data Analysis, New York: John Wiley.
Hoaglin, David C., Mosteller F., Tukey J. (1985) Exploring Data Tables, Trends
and Shapes, New York: John Wiley.
Holden, Darryl and Perman, Roger (1994) 'Unit Roots and Cointegration for the
Applied Economist', in B. Bhaskara Rao (ed.) Cointegration far the Applied
Economist, Oxford: Basil Blackwell.
Hopwood, A. (1984) 'Accounting and the Pursuit of Efficiency', in A. Hopwood
and C. Tomkins, lssues in Public Sector Accounting, Oxford: Phillip Allan,
167-87.
Huber, Peter J. (1981) Robust Statistics, New York: John Wiley.
Kennedy, Peter (1992) A Cuide to Econometrics, Oxford: Blackwell.
Khan, M.S. and Reinhart, C.M. (1990) 'Prívate Investment and Economic Growth
in Developing Countries', World Development 18: 19-28.
Kmenta, Jan (1986) Elements of Econometrics, New York: Macmillan.
Krishnaji, N. (1992) 'The Demand Constraint: A Note on the Role of Foodgrain
Prices and Income Inequality', in N. Krishnaji, Pauperising Agriculture: Studies
in Agrarian Change and Demographic Structure, Sameeska Trust, Bombay:
Oxford University Press.
Leamer, E.E. (1978) Specification Searches: Ad hoc lnference with Non-experi-
mental Data, New York: John Wiley.
Leamer, E.E. (1983) 'Let's Take the Con Out of Econometrics', American
Economic Review, 23 (1): 31-43.
Levine, J.H. (1993) Exceptions are the Rule: An lnquiry into Methods in Social
Sciences, Boulder, CO: Westview Press.
Lucas, R.E. (1976) 'Econometric Policy Evaluation: A Critique', in K. Brunner
and A.H. Meltzer (eds), The Phillips Curve and Labour Markets, supplement
to Journal of Monetary Economics 1: 19-46.
MacKie-Mason, J.K. (1992) 'Econometric Software: A User's View', Journal of
Economic Perspectives 6 (4): 165-87.
McKinnon, R.I. (1973) Money and Capital in a Developing Economy, Washington,
DC: Brookings lnstitution.
Maddala, G.S. (1988) lntroduction to Econometrics, Englewood Cliffs: Prentice
Hall.
Maddala, G.S. (1992) lntroduction to Econometrics, New York: Macmillan.
Miller, R.W. (1987) Fact and Method: Explanation, Confirmation and Reality in
the Natural and the Social Sciences, Princeton, NJ: Princeton University Press.
Moore, D.S. and McCabe, G.P. (1989) lntroduction to the Practice of Statistics,
New York: Freeman.
Morgan, M.S. (1990) The History of Econometric Ideas, Cambridge: Cambridge
University Press.
Mosley, Paul, Harrigan, Jane and Toye, John (1991) Aid and Power: The World
Bank and Policy-based Lending, 2 vols, London: Routledge.
Mosteller, Frederick and Tukey, John W. (1977) Data Analysis and Regression: A
Second Course in Statistics, Reading, MA: Addison-Wesley.
Myers R.H. (1990) Classical and Modern Regression with Applications, 2nd edn,
Boston, MA: PWS-Kent.
484 Econometrics far developing countries
Pelto, Pertti J. and Pelto, Gretel H. (1978) Anthropological Research: The Structure
of Inquiry, Cambridge: Cambridge University Press.
Phillips, P.C.B. and Ouliaris S. (1990) 'Asymptotic Properties of Residual Based
Tests for Cointegration', Econometrica 58 (1): 165-93.
Rao, B. Bhaskara (ed.) (1994) Cointegration far the Applied Economist, London:
Macmillan.
Rawlings, John O. (1988) Applied Regression Analysis: A Research Too/, Pacific
Grove, CA: Woodsworth & Brooks/Cole.
Ridde!l, Roger (1987) Foreign Aid Reconsidered, London: James Curry.
Rosenberger James L., and Gasko, Miriam (1983) 'Comparing Location
Estimators: Trimmed Means, Medians, and Trimean', in David C. Hoaglin, F.
Mosteller and J. Tukey Understanding Robust and Exploratory Data Analysis,
New York: John Wiley, pp. 297-338.
Ross, J.A., Rich, M., Molzan, J.P. and Pensak, M. (1988) Family Planning and
Child Survival, 100 Developing Countries, Centre for Population and Family
Health, New York: Columbia University.
Sapsford, David (1985) 'The Statistical Debate on the Net Barter Terms of Trade
Between Primary Commodities and Manufactures: A Comment and Sorne
Additional Evidence', Economic Journal 95: 781-8.
Seers, D. (1976) 'The Political Economy of National Accounting', in A. Caimcross
and M. Pur (eds), Employment, Income Distribution and Development Strategy,
London: Macmillan.
Sen, A.K. (1985) 'Women, Technology and Sexual Divisions', Trade and
Development (UNCTAD), 6.
Sen, A.K. and Sengupta, S. (1983) 'Malnutrition of Rural Indian Children and the
Sex Bias', Economic and Political Weekly 18.
Sen, Gita (1993) 'Paths of Fertility Decline: A Cross-country Analysis', in Pranab
Bardhan, Mrinal Datta-Chauduri and T.N. Krishnan, Development and Change,
Bombay: Oxford University Press.
Shaw, E. (1973) Financia/ Deepening in Economic Development, New York:
Oxford University Press.
Sims, C. (1980) 'Macroeconomics and Reality', Econometrica 48: 1-48.
Snedecor, George W. and Cochran, William G. (1989) Statistical Methods, New
Delhi: Affiliated East-West Press.
Spanos, Aris (1986) Statistical Foundations of Econometric Modelling, Cambridge:
Cambridge University Press.
Spanos, A. (1990) 'Towards a Unifying Methodological Framework for Econo-
metric Modelling', in C.W.J. Granger (ed.), Modelling Economic Series, Oxford:
Clarendon Press.
Sproas, J. (1980) 'The Statistical Debate on the Net Barter Terms of Trade
Between Primary Commodities and Manufactures', Economic Journal 90:
107-28.
Stigler, S.M. (1986) The History of Statistics: The Measurement of Uncertainty
befare 1900, Cambridge MA: Belknap Press.
Tukey, J.W. (1977) Exploratory Data Analysis, Reading, MA: Addison-Wesley.
Wheeler, E.F. (1984) 'Intra Household Food Allocation: a Review of Evidence',
paper presented at meeting on 'The Sharing of Food', Bad Homborg, London:
London School of Hygiene and Tropical Medicine. mimeo.
White, Halbert (1980) 'A Heteroscedasticity Consistent Covariance Matrix
Estimator and a Direct Test for Heteroscedasticity', Econometrica 48: 817-38.
White, Howard (1992) 'What do we Know About Aid's Macroeconomic Impact?',
Journal of International Development 4: 121-37.
Working, H. (1943) 'Statistical Laws of Family Expenditure', Journal of the
American Statistical Association 38: 43-56.
Index

absolute residuals (plots) 254-5, 267, average economic regression 24, 29-32
273-6 average propensity to consume 406,
actual values 403, 405, 444, 445 407, 409-11
added-variable plot 169-70, 183--4 averages (kinds of) 45-50
ad hoc modifications (model
specification) 31-2, 33, 36, 276 Banerjee, Anindya 41, 352
African economies (GNP per capita) Barten, A. P. 112, 158
77-8, 79, 80, 86, 98 Bartlett's test 256-9, 263--4, 267, 270,
age of workers, wages and 277
(heteroscedasticity) 253-6, 261-6 bell-shaped curve 27, 28, 46-8, 49, 65
aid 455; Griffin's model 38, 208, Belize (population data) 372-5, 384,
209-12, 214, 219, 222, 225, 227-8 458
Aldrich, John H. 325 benchmarks, dummy variables and 284
algebra of stationarity 343-6 best linear unbiased estimator
Altman, J. C. 29 (BLUE) 418; autocorrelation and
analysing cross-section data: see cross- misspecification 366, 370, 387; least
section data squares estimators 75, 78, 111, 119,
antilogarithm (inverse transformation) 164, 185, 251, 269, 272, 370;
103-5, 195, 285, 311-12, 323 modelling an average 45, 53-8, 61
applied research 2-3; model bias: assessing direction of 214-15;
specification and 23--43 omitted variable 208, 209-19,
AR(l) process 338--48, 351-2, 367-9, 269-70
380-1, 385, 387, 391, 398 binomial distribution 289-90
arithmetic progression 158 birth rate 4-5, 7-16, 19, 187-91, 456
association: between categorical bivariate analysis 51, 125, 313-16, 399,
variable 287-92; causality and 406
118-19 bivariate normal distribution 124, 126,
augmented Dickey-Fuller test 352, 361
399, 401, 402-3, 406 see bivariate relations 9
Dickey-Fuller test BLUE see best linear unbiased
autocorrelation 132--4, 229, 335, 348, estimator (BLUE)
352; misspecification and (time- Bowley's resistant coefficient 93, 106,
series data) 366-92 132
autoregressive distributed lag (ADL) Box, G. E. P. 40
408 box plots: data analysis (simple
autoregressive process 338--48, 351-2, regression) 132, 142-3, 146, 150;
367-9, 380-1, 385, 387, 391, 398 data transformation( outliers and
auxiliary regressions 169, 176, 182, skewness) 86-93; logit
185, 213, 216 transformation 305-6;
average (modelling) 44-74 transformations to stationarity 357
486 Index
business cycle 345 collinearity 198, 218-19; perfect (of
regressors) 177, 178-9
categorical data (counts and competing proxy variables (regression
measurements) (cross-section data) coefficients) 203-5, 218, 219
279-301 computing the median/quartiles 84-7
categorical variable: association conditional bounds on regression
between (contingency tables) coefficients 202-3
287-92; dependent 302-32; multiple conditional effects plot 328-9
regression on 295-8; regression on conditional mean 122, 138, 265
(using dummy variables) 280-7 confidence intervals 62-6, 96-7, 105,
causality: association versus 118-19; 121-3, 383-4
econometric tests of 415, 425-8, 434 constant mean 361-2
Chambers, John M. 28 constant returns to scale 229, 230
Charemza, Wojciech W. 34 constant term 327; dummies and
checking: assumptions of regression 284-5, 314
model 187-92; model assumptions consumption, income and 406--7
(testing zero restrictions) 226-9; for consumption function 433;
normality in data 91-5 cointegrating 405-6, 407, 409-11,
Chenery, Hollis B. 224 456; Costa Rica 402-5, 456;
chi-square distribution 74, 95, 258 estimating (in simultaneous system)
chi-square test 6, 324-6, 330--1; 445-8; Indonesia 445-8
contingency (of independence) contingency chi square test of
290--2 independence 290-2
child mortality 202-3, 204 contingency tables: association
China (household size/income data) between categorical variables
98-100, 101, 456 287-92; logit modelling with 307-13
Chow test 233, 245, 246--7; second test correlation coefficient 118, 176, 180-4,
236-7 214
classical linear regression model correlograms 379-83
115-16, 120-4 Costa Rican consumption function
Cobb-Douglas production function 402-5, 456
229 counts (cross-section data) 279-301
Cochran, William G. 96 covariance, derivation of (AR(l)
Cochrane-Orcutt procedure 387-90, model) 391
391 Cramer's V statistic 292-4, 299
coefficient of determination 5, 6-7, 14, critical values 292, 352, 354-5, 385,
118, 171-2, 179-80, 185 386, 399, 402, 474-5, 480
coefficient of kurtosis 81-3, 87 crop production function 371-2,
coefficient matrix (rank and arder 375-6, 380, 382-4, 386, 388-90
conditions) 433 cross-correlation 448, 449
coefficient matrix (supply/demand cross-section data: autocorrelation in
model) 431-2 376--9; categorical data 279-301;
coefficient of partial correlation 181-2, heteroscedasticity 251-78; logit
193-4 transformation/modelling 302-32;
coefficient of skewness 81-3, 87, 93, time-series data and 34, 39-42
106 crowding-in (investment) 224, 225,
coefficient of variation 81, 89-90 244-5
cointegrating consumption function crowding-out (investment) 225, 244-5
405-6, 407, 409-11, 456 cubic powers 82
cointegrating regression DW (CRDW)
402, 403, 406 Dasgupta, P. 4-5
cointegration 41, 335; error correction data: exploratory check for normality
model and 393-412 91-5; exploring (graphical methods)
Collier, P. 462 7-13; mining 3, 32; modelling with
References 487
transformed 102-5; non-normality in DFBETA statistic 145-8, 190, 192, 228
(detecting) 90-7; pooling (tests of Diaconis, P. 32
parameter stability) 231-4; role (in diagnostic graphs (in logit regression)
model specification ) 29-39; sets 327-31
18-20, 455-62; table (collapsing) diagnostic plots (heteroscedasticity)
308; time dimension (in model 252-6
specification) 39-42 diagnostic testing (initial model
data analysis (aims/approaches of specification/estimation) 4-7
study) 1-20 dichotomous dependent variables
data analysis (foundations of): model 302-3, 306-8, 310, 312, 313
specification and applied research Dickey-Fuller test 352, 399, 401-4,
23-43; modelling an average 44-74; 406, 412, 480
outliers and skewness (data trans- difference stationary process (DSP)
formations) 75-107 349-51, 365
data analysis, regression and: model differences/differencing 360, 362-3,
selection and misspecification 388
208-48; partía! regression 163-207; Direction of Trade Statistics
simple regression 111-62 (Pakistan) 459
data generation process (DGP) 212, domestic savings 38, 208, 209-12, 233-4
366, 368-9, 394, 405 double-log transformation 149-53, 265,
data transformations: to eliminate 273, 275, 276
skewness 97-105, outliers and Ducpetiaux, E. 112
skewness 75-107 dummy variables 313-14, 317, 322-3;
Deadman, Derek F. 34 parameter instability 237-46;
decision tree (stationarity testing) regression on categorical variable
352-4, 399, 401 280-7, 296, 298
degrees of freedom 121, 258, 325; Durbin-Watson statistic 19, 338, 348,
model selection 216-17; modelling 366, 379, 384-91, 398-404, 406, 412,
an average 66, 69, 74; outliers and 470, 472
skewness 82, 95; partial regression
186, 194 econometrics/econometric modelling 2,
demand: function (money illusion in) 23-4, 29-32; tests of causality 415,
166-7; for labour (scatter plot) 425-8, 434; time-series data and
126-8; for manufactured goods 39-42
(India) 165-73, 194-7 Eisenhart, C. 475
demand and supply model 441, 460; elasticities 134, 149, 159, 410-11;
identification in 428-30; simultaneity constant elasticity function 165;
bias in 417-22 income 166, 196, 205, 239; price 166,
Demaris, Alfred 307 415, 418, 423, 441; supply 418, 423,
density function 58, 59 441, 444
dependent variables 7-9, 219; Emerson, John D. 86, 173
categorical (logit transformation) empírica! research: encompassing
302-32; dichotomous 302-3, 306-8, approach 211-12; exploratory data
310, 312, 313; lagged 366, 370, 386; analysis 35-7
numerical 280-7; partía! regression encompassing approach (in empírica!
163-5, 168, 171-2, 179, 189-90; simple research) 211-12
regression 111, 118, 153-4, 156, 168 endogenous variables 415-18, 420-1,
determination, coefficient of 5, 6-7, 425, 428, 430, 444
14, 118, 171-2, 179-80, 185 energy consumption 150-6,
deterministic time trend 344, 347, 352, Engel's law 24-5, 112, 158
355, 360 Engle, Robert F. 402, 404
detrending 360 Enos, John 209
devaluation, response of Pakistaní error correction model, cointegration
exports to 399-402 and 41, 387, 393-412
488 Index
error terms continued fertility data 4-5, 15, 198-9, 200,
error correction term 407-8 457-8
error terms 76, 252; logit transform fitted values 403, 405, 444, 445
315-16, 327; misspecification and five-number summary 86, 87
autocorrelation 366-7, 379-80; tlow variables 39
modelling an average 52-3, 67; simple focus variable 38
regression 111, 114-16, 121, 150; food: expenditure 112-13, 158-9, 271,
simultaneous equation models 417, 273-6; prices (India) 165-73, 194-7,
438; spurious regression 338, 340--1, 230--1, 458
344, 346-7 foreign aid 38, 208, 209-12
error variance 6, 25, 186-7 fourth root transformation (per capita
errors: in hypothesis testing 67-8; household income) 101-2
standard see standard errors fragility analysis 37-8; regression
estimation: in logit regression 320-7, coefficients 198-205
of population variance 65; of Friedman, Milton 338
simultaneous equation models Fry, M. J. 224
437-53 full information estimation techniques
estimator of error variance 186-7 (simultaneous equation) 445,
exchange rates (Pakistan) 399-402 448-51
exogeneity 416, 425, 427-8 Fuller, W. A. 480; see also
exogenous variables 415-21, 424, 425, Dickey-Fuller test
428-9, 431, 442-4
expenditure, recurrent 142-3, 144, 145; Gasko, Miriam 45
Tanzania 128-37, 239-42 Gatt, J. 420, 458
explained sum of squares (ESS) Gauss, Karl Friedrich 78
117-18, 168-9, 171-2, 179-80 Gauss-Markov theorem 119, 163, 185
explanatory variables 111; categorical gender differences (life expectancy)
280-7; logit modelling 310--11, 312, 87-90, 91-3, 94
317; partial regression 163-4, general-to-specific modelling 33-5,
168-72, 176-7, 198 136, 165, 198
exploratory band regression 113, generalised least squares (GLS)
125-8, 131, 241 estimation 448, 449
exploratory check for normality in generated DSP series 365
data 91-5 generated TSP series 365
exploratory data analysis (EDA) 44, 'genuine' autocorrelation 387-90
125; model specification 35-7, 39; geometric progression 158
outliers and skewness 75, 80, 86-7, Giere, Ronald N. 26
90, 100; partial regression plot 164, Gilbert, Christopher 34
169-70, 183-4, 201 Glejser's test 256, 261-4, 267, 270, 277
exports: Malta 212, 420--1, 423, 426, GNP per capita 4-5, 6, 14-16, 187-91;
458; Pakistan 399-402, 459 African economies 77-8, 79, 80, 86,
extreme bounds analysis (fragility) 98; energy consumption and 150--3;
37-8, 198-205 graphical methods of data analysis
7-13; life expectancy and 87-90,
F-statistic 6, 225-6, 261; distribution 91-3, 94, 157-8, 232
221, 466-7, 476-9; t-test and Goldberger, Arthur S. 44, 124, 316
(comparison) 208-9, 222-3; tests Goldfeld-Quandt test 6, 14, 250,
165, 192, 195, 198, 228-9, 233, 259-61, 263-4, 267, 270, 277
235-6, 245, 355, 363, 420, 425-7; Gould, Stephen Jay 3
zero restrictions 208-9, 220--3 Granger, C. W. J. 7, 23, 32, 37, 41,
family planning variable 202-3 402, 404
Farebrother, R. W. 470, 472 Granger causality test 415, 425-6,
fat tails 100 427-8, 434
female literacy 202-3, 204 graphical methods 3, 4, 7-13
Index 489
graphics: in logit regression 327-31; identification: problem (simultaneity
regression with (checking model bias) 428-34; in supply and demand
assumptions) 124-35 model 428-30
Gregory, C. A. 29 income: consumption and 406--7; data
Griffin, Keith (aid model) 38, 208, 98-101, 258-9, 261-7, 456; house-
209-12, 214, 219, 222, 225, 227-8 hold 98-103; life expectancy and
Grilli, E. 234, 462 377-8; variable 167-9
grouped data 271-2, 326--7 incorrect functional form 372-5
growth rates, logged differences and 360 incremental capital-output ratio
Gujarati, D. 463 (ICOR) 210
Gupta, K. 211 independe11ce: contingency chi-square
test 290--2; stochastic 287-8, 289,
Haavelmo-Cowles research 291
programme 30 independent variables 118, 154, 190,
Hamilton, Lawrence C. 7, 36, 44, 125, 303
126, 144, 148, 190,325,327,329,406 India: food/total expenditure 268, 269,
Harriss, B. 90 457; food prices/manufacturing
Harrod-Domar equation 209, 210 demand 165-73, 194-7, 230--1, 458;
Hartley, H. O. 465 wages/age of workers 253-6, 457;
hat statistic 115, 123, 142, 143-5, 147, wages and education/gender 269-70,
190, 192 279-98, 456--7
Hausman specification test 418-21, indirect least squares (ILS) 439-42,
424, 428, 434, 445 443, 444-7
heavy tails 80, 81, 82-3, 105; checking individual case data, logit regression
for (normality assumption) 93-4; with 321--6
mean/median ( estimators) 96--7 Indonesia: consumption function
Heckman, J. J. 30, 35 445-8; national accounts data 451,
Hendry, D. 32, 34 457
heteroscedastic standard errors infant mortality 4-5, 6, 7-16, 187-91
(HCSEs) 123, 251, 270--7 inflation, money supply and 338-9,
heteroscedasticity 14, 123, 126, 129, 348-9, 361
131, 149, 150, 152, 190, 229, 327; influence, outliers and leverage 137-48
dealing with 251-78; in linear influential points (DFBETA statistics)
probability model 315-16 145-6
histograms 7-9, 11 initial model specification/estimation
Hoaglin, David C. 85, 87, 94, 125, 173 (with diagnostic testing) 4-7
Holden, Darryl 352, 354, 406 instrumental variable estimation
homoscedasticity 6--7, 115, 119, 123; 417-19, 420, 423-4, 442-6, 447
transformations towards 261, 264-70 integrated variables 351
household: expenditure 112-13, 158-9, interaction effect: logit modelling with
190, 271, 273-6; income (China) contingency tables 311-12; partía!
98-100, 101-2, 103; size (China) association 293-4, 298
98-100, 101, 456 intercept dummies 237-41, 243, 244-6,
Huber, Peter J. 144 298
Human Development Index 187-91, interest rate 415
442 International Financia! Statistics
Human Development Report (UNDP) (IMF) 456, 457, 459, 461-2
458, 460 interpreting (error correction model)
hypothesis testing 27, 39; classical 407-12
normal linear regression model 123; interquartile range (IQR) 85, 87, 89,
in logit regression 320--7; model 90, 91, 93-4, 358-9
specification 208-48, modelling an intuitive approach (cointegration)
average 44-74; testing downwards 394-9
33-5 invariance of series variance 337-8
490 Index
inverse transformational gender and 87-90, 91-3, 94
(antilogarithm) 103-5, 195, 285, limited information estimation 445,
311-12, 323 448
investment: crowding-in 224, 225, linear combination between regression
244--5; crowding-out 225, 244--5; Sri coefficients, testing for 195-7
Lankan investment function 223-5, linear infiuence of income variable,
243-6 removing (partial regression) 168-9
irrelevant variables (omitted variable linear probability model 313-20
bias) 213-14, 218-19 linear regression 372-4; least squares
IS curve 415 principie and 111, 114-20; model
(inference from) 120-4; model
Jarque-Bera test (skewness-kurtosis) (partial regression) 184--92
6, 14, 94--5, 100, 132, 259, 276, 296 linear restrictions, non-zero 229-31
Jenkins, G. M. 40 linearity, transformation towards
Johansen method (cointegration test) 148-59
399, 406 LM curve 415
logarithmic transformation 14, 358, 362;
k-variable case 178-9, 182-3, 185 antilogarithm 103-5, 195, 285,
Kennedy, Peter 32, 34, 37 311-12, 323; double-log trans-
Keynesian macro model 432-3 formation 149-53, 265, 273, 275, 276;
Keynesian model 415, 434--5, 438-9, to eliminate skewness 98-104; semi
451 logarithmic transformation 153-9;
Khan, M.S. 231 versatility of (heteroscedasticity) 259,
Kmenta, Jan 322, 327 266, 268-9, 273 logged differences
Krishnaji, N. 165-73, 196-7, 205, and rates of growth 360
230--1, 458 logged real exchange rates (Pakistan)
kurtosis 81-3, 87, 90; Jarque-Bera test 399-400
6, 14, 94-5, 100, 132, 259, 276, 296 logit modelling with contingency
tables 289, 307-13
labour demand 46-8, 50, 53, 55-6, 60, logit regression: estimation and
62, 458-9 hypothesis testing in 320--7; graphics
ladder of power transformation 9, and residual analysis 327-31; linear
100--2, 105, 259, 356-8 probability model and 313-20; with
lagged dependent variable, auto- saturated model 312-13
correlation with 366, 370, 386 logit transformation (cross-section
Landau, L. 211 data) 302-32
Leamer, E. E. 27, 32, 34, 37-8, 198, lognormal variable 150
425 Lucas, R. E. (Lucas critique) 428
least squares estimation 13, 116-17,
120, 124--5, 163; as BLUE 75, 78, McCabe, G.P. 138
111, 119, 164, 185, 251, 269, 272, McKinnon, R. l. 224, 352
370 Madarassay, A. 461
least squares line 170, 173-80 Maddala, G. S. 124, 428
least squares principie 137; concept of Malinowski, Bronislaw 29
resistance and 76-80; indirect least Malta (exports) 212, 420-1, 423, 426,
squares 439-42, 443, 444-7; linear 458
regression and 111, 114--20, 174--5; manufactured goods, demand for
three-stage least squares 448, 450-1; (India) 165-73, 194--7, 230--1, 458
two-stage least squares 420--1, 424, Maputo harbour: casual labour 46-8,
442-5, 447, 451 50, 55, 60, 62-4, 67-70, 83, 94, 126,
Legendre, Adrien Marie 78 149, 317-20, 458-9; overtime
leverage, outliers and infiuence 137-48 payments 48-9, 53
Levine, J. H. 3 marginal propensity to consume 407,
life expectancy 157-8, 232, 377-8, 458; 409-11
Index 491
mathematical properties: of least money supply 363, 415; infiation and
squares regression line 117-18, 338-9, 348-9, 361
174-5; of sample mean 76-8 Moore, D. S. 138
maximum bounds (regression Morgan, Mary S. 30
coefficient) 201-3 Mosley, Paul 247-8
maximum likelihood principie 321-2, Mosteller, Frederick 35, 37, 60, 79,
324-5; normality and 58-61, 112, 204
65; normality assumption 111, multi-equation model 437-8
119-20 multicollinearity 182, 186, 198, 202,
mean 73; based statistics 78-91, 133; 216, 217
conditional 122, 138, 265; estimator multiple regression 6, 14, 16, 111-12;
(heavy tails) 96-7; median versus on categorical variables 295-8;
60-1, 79-80, 91-3, 96-7, 105; in coefficients (interpreting) 163-207;
modelling an average 45-50; model selection and misspecification
stabilising 360-2; see a/so sample in 208-48; partial regression and
mean; zero mean 170-1, 180-1; t-test in 192-8, 219,
measurements (cross-section data) 239-42
279-301 multivariate analysis 16, 399, 406;
median: computing 84-7; estimator see a/so multiple regression
(heavy tails) 96-7; interquartile multiway contingency tables 303, 307
range and 358-9; mean versus 60-1, Myers, R. H. 138, 241
79-80, 91-3, 96-7, 105; in modelling
an average 45-50 Nelson, Forrest D. 325
Miller, R. W. 30, 31 Newbold, P. 41
minimum bound of (regression nominal categorical variable 282
coefficient) 201-3 non-linearity 127, 131, 149, 150, 152;
minimum variance property of sample eliminating skewness 97-105;
mean 56-8 heteroscedasticity 126, 261, 262
misspecification: autocorrelation and non-normality in data (detecting) 90-7
(time-series data) 366-92; bias non-stationarity 335-8, 345, 348-50;
(single equation estimation) 415-36; cointegration and 393, 396, 404,
model selection and (in multiple 406-7; testing for (unit roots) 351-6,
regression) 208-48 386-7, 402
mode 45-50 non-trade variables 308-9
model assumptions (regression with non-zero determinants (rank and
graphics) 124-35 order conditions) 431, 433
model closure 420 non-zero linear restrictions ( testing)
model estimation/testing: modelling an 229-31
average 44-74; see also statistical normal distribution, inference from
inference sample of 61-71
model selection 208-48 normality: in data ( exploratory check)
model specification 4-7; and applied 91-5; maximum likelihood principie
research 23-43; role of data 29-39; 58-61; skewness-kurtosis test 6, 14,
statistical inference and 24-9 94-5
modelling: an average 44-74; cross- normality assumption 6; linear
section data 302-32; a random walk regression model 120-4; maximum
338-43; simple regression 112-14; likelihood and 111, 119-20; model-
with transformed data 102-5 ling an average 17, 45, 58-61;
modern approaches to model validity (detection procedures) 90-7
specification (role of data) 32-9 null hypothesis 6, 95, 135, 325, 354-5;
moments of distribution (skewness cointegration 401-2, 404, 406;
and kurtosis) 106 heteroscedasticity 14, 257-8;
money illusion (in demand function) misspecification and autocorrelation
166-7, 196, 231 369, 383-6; misspecification bias
492 Index
418-19, 421, 426-7; modelling an Pelto, Gretel, H. 31
average 66-9, 70; multiple regres- Pelto, Pertti J. 31
sion 222, 225, 229, 233, 236; partial perfect collinearity of regressors 177,
regression 194, 196 178-9
numerical dependent variable 280-7 Perman, Roger 352, 406
numerical summaries 125, 126, 132-3 permanent income hypothesis 405
Peru (database) 459
odds ratio 310-12, 323-4 Pfefferman, G. 461
omitted variables 262; bias 208-19, Phillips, P. C. B. 402, 404
269-70; as cause of autocorrelation Plosser-Scwert-White differencing test
375-6; Hausman test 419-22; 362-3
heteroscedastic residuals due to 262; pooling data(tests of parameter
Plosser-Scwert-White test 362-3 stability) 231-4
one-tailed test 68, 69 population: data (Belize) 372-5, 384,
order-based sample statistics 80-90, 458; distributions 55-6, 60-1; mean
91, 133 49, 53-5, 56, 57, 59, 60-4, 67, 70,
order condition 430-4 102-3; regression 112, 114, 115, 125,
order of integration (integrated 208, 270-1, 372-5, 384, 455; variance
variables) 351 ( estimating) 65, 66
ordinal categorical variable 281 power transformation 152, 159, 265,
ordinary least squares 148; auto- 360; ladder of 9, 100-2, 105, 259,
correlation 366, 386, 388-9; 356-8
cointegration 394, 400-1, 407; logit Prais-Winsten transformation 388
transformation 316, 319, 326-7; Prebisch, Raul 234
misspecification bias 415-24, 428-9; predicted values (plot) 253-5, 267,
multiple regression 212, 229; 273-4
simultaneous equation models predictive failure 236-7
437-8, 440-1, 443-6, 449-51; prices of food (India) 165-73, 194-7,
spurious regressions 345, 347-8, 350, 230-1, 458
362 private investment (Sri Lanka) 223-5,
origin, regression through 136-7 243-6
orthogonality of regressors 177, 178-9, probability distributions 45
198, 213-14 probability limits (plims) 423-4
Ouliaris, S. 402, 404 probability theory 25, 61-3
outliers 13, 14; definition 87; detecting production function 230, 460;
87-90; leverage and influence Cobb-Douglas 229; crops 371-2,
137-48; skewness and data 375-6, 380, 382-4, 386, 388-90
transformations 75-107 proxy variables, competing (regression
over-identification (indirect least coefficients) 203-5, 218, 219
squares) 441-3, 446-7 pseudo-R2 325-6
overshooting 409 pseudo standard deviation 94, 132
overtime payments (Maputo harbour) public investment (Sri Lanka) 223-5,
48-9 243-6

Pakistan ( exports) 399-402, 459 quadratic term, testing for 197-8


parameter instability (use of dummy qualitative dependant variable 312,
variables) 237-46 313, 319, 320-1
parameter stability (tests of) 231-7 quantitative dependant variable 293-4,
partial association 287, 293-4 298
partial correlation 180-4, 193-4 quartiles, computing 84-7
partial regression 14, 163-207, 227
Pearson, E. S. 354, 465 R 2 : definition 118; in partial regression
Pearson residuals 329-30, 332 172; in logit regression: see pseudo-
Pearson's Chi-square 329-31, 332 R2; relation to simple correlation
Index 493
coefficient 118; testing restrictions regressor (dropped from model)
with 222 222-3
random series, independently Reinhart, C. M. 231
generated ( significant regions) 346-8 relative efficiency of mean/median
random walk 40, 41, 353--4, 356; 60--1
spurious regression and 338--49, 393, replicated data 326
394, 396-8 RESET 261, 262
rank and arder conditions 430--4 residual analysis in logit regression
Rao, B. Bhaskara 406 327-31
rates of growth, logged differences residual autocorrelation 352
and 360 residual sum of squares (RSS) 117-18,
Rawlings, John O. 265 120, 168-9, 171-2, 179-80, 195,
real exchange rate (Pakistan) 399--402 216-17, 220--2, 233, 235-7, 239--40,
real manufacturing GDP (Tanzania) 354-5
50-1 residuals 36, 44, 76-7, 366, 370, 372--4,
recruitment of labour (scatter plot) 380, 383; absolute 254-5, 267, 273-6;
126-8 heteroscedastic 252-3, 261-2, 273-6;
recurrent expenditure 142-3, 144, 145; plots 128-35, 138--41, 190, 252-5,
Tanzania 128-37, 239--40, 241-2, 462 369, 375-7, 379, 396-8; squared 254,
recurrent revenue 142, 144, 145; 255, 261, 276, 445; studentised
Tanzania 128-37, 239--40, 241-2, 462 142-3, 147, 190, 192, 228, 238--41
recursive models (simultaneous resistance concept, least squares
equation) 416, 437-9 principle and 76-80
regression: on categorical variables resistant summaries 78-80, 91, 125,
280--7; coefficients 195-205, 215-17; 137
cross-section data 302-32; curve 15; restricting slope coefficients to zero
with graphics 124-35; line 175, 225-6: see also restrictions-testing
318-20; through the origin 136-7; Riddell, Roger 211
seemingly unrelated 438, 448-50; robust summaries 137
significant (independently generated Rosenberger, James L. 45
random series) 346-8 Ross, J. A 457
regression (with time-series data): runs test 369, 383--4, 474-5
cointegration and error correction
model 393--412; misspecification and sample mean: as best linear unbiased
autocorrelation 366-92; trends estimator 45, 53-8, 61, 75; checking
335-65 skewness 91, 103, 104-5; distribution
regression analysis 1-2, 44; applied of 61-6, 70; as estimator of
research (model specification) population mean 53-8, 60--4, 70, 96,
23--43; approach adopted (an 105; least squares principie 76-80;
example) 3-16; categorical data mathematical properties of 76-8;
279-301; multiple regression see minimum variance property of 56-8;
multiple regression; partial unbiasedness of 55-6
regression 14, 163-207, 227; simple sample median 91, 96-7
regression 14, 111-62, 167-8 sample multiple regression line 173-80
regression and data analysis: model sample statistics (mean-based and
selection and misspecification order-based) 80-90
208--48; partial regression 163-207; sampling distributions 55-6, 60-1,
simple regression 111--62 73--4, 121, 185-6
regression model: checking sampling procedure 26-7, 54
assumptions of 187-92; with Sapsford, David 234
transformed explanatory variables saturated model, logit regression with
13-16 312-13, 327
regression specific error test (RESET) Savin, N. E. 470, 472
261, 262 savings 233--4, 455; Griffin's model 38,
494 Index
208, 209-12, 214, 219, 222, 225, spatial correlation 377
227-'d specific-to-general modelling 29-32
scatter plots 15, 126-31, 150-1, 167-8, specification bounds on regression
226-7, 232, 403-4; matrix 9-10, coefficients 199-202
12-13, 14, 187-90 Sproas, J. 234
seemingly unrelated regressions 438, spurious correlations 40-1, 362
448-50, 461 spurious regressions 387; random
semi-logarithmic transformations walks and 338-49; time-series data
153-9 335-65, 393-5, 396-8
Sen, A.K. 90, 199 squared residuals 254, 255, 261, 276,
Sengupta, S. 90 445
sensitivity analysis 37-8; regression Sri Lanka; 161, 461; investment
coefficients 198-205 function 223-5, 243-6
serial correlation 132-4, 229, 335, 348, stabilising the mean 360-2
352; misspecification and (time- stabilising variances 356-9
series data) 366-92 standard deviation 69, 70, 78, 81-2,
sex bias against women 87-90 89-90, 93-4, 96
Shaw, E. 224 shrinking effect of standard errors 55, 120, 215-17;
logarithms 98 heteroscedasticity 123, 251, 270-7
Siegel, Sidney 475 standard normal distribution 61-2
significance ( statistical and standard sampling distributions 73-4
substantive) 70-1 stationarity: algebra of 343-6;
significant regressions of cointegration and 393, 394, 399, 406;
independently generated random testing (decision tree) 352-4, 399,
series 346-8 401; transformations to (time-series
simple regression 14, 111-62; ignoring data) 335-65
income variable 167-8 statistical inference 24-9, 61-71,
Sims, C. 345 120-4
Sims test (for causality) 415, 426, 427 statistical model assumptions
simulation of error correction models (regression with graphics) 124-35
409-10 statistical properties (linear regression
simultaneity bias 415-25, 438, 442 model) 185
simultaneous equation models 2; statistical significance 70-1
estimation of 437-53; statistical study ( elements of) 26-7
misspecification bias 415-36 statistical tables 463-'dO
simultaneous systems (terminology) statistical tests of heteroscedasticity
416 256-64
Singer, l. M. 234 Stigler, S. M. 78
single equation estimation 415-36 stochastic independence 287-8, 289,
skewness 269; checking for 91-3; data 291
transformation to eliminate 9, stochastic trend 347
97-105; Jarque-Bera test 6, 14, 94-5, stochastic variables 25, 27, 123-4, 361
100, 132, 259, 276, 296; outliers and stock variables 39
data transformations 75-107 Strenio, Judith 86
slope coefficients 5, 120, 135, 154, 156, Strout, William 224
159, 164, 166, 168-9, 171, 176, structural adjustment policies 308, 310,
178-9, 195-6, 213-15, 237-8, 240, 311, 415
314; restricting to zero 225-6 structural equations 416, 420, 438, 444
slope dummies 237, 241-6 studentised residuals 142-3, 147, 190,
Slutsky's theorem 422, 423 192, 228; intercept dummy
Snedecor, George W. 96 interpretation of 238-41
Social Indicators of Development substantive significance 70-1
460 sums of squares, decomposing (across
Spanos, A. 34, 408 regressions) 171-2
Index 495
superfluous variables 214; partial time-series data 416; autocorrelation
regression 177, 178, 179 and 132-4; cointegration and error
supply and demand model 441, 460; correction model 393-412; cross-
coefficient matrix 431-2; section data and 34, 39-42;
identification in 428-30; simultaneity misspecification and autocorrelation
bias in 417-22 366-92; parameter stability 234-6;
supply response 441-2, 444-5 trends ( spurious regressions) 335-65
survey data 2 total fertility rate 198-9, 200
Swed, Frieda S. 475 total sum of squares (TSS) 117-18,
sweeping out (partial regression) 120, 168-9, 171-2, 179-80
172-3, 177, 193-4 trade/non-trade variable 308-9
symmetry, approximate (checking traditional modelling (role of data)
skewness) 91-3, 96 29-32
t-statistic/test 13-14, 69, 142, 219, transformations: double-log 149-53,
324; cointegration 400-4, 407; 265, 273, 275, 276; to eliminate
distribution 65-7, 74, 82, 121, 123, heteroscedasticity 251-2, 265-9;
465; F-test and (comparison) 208-9, towards homoscedasticity 264-70;
222-3; multiple regression 192-8, inverse (antilogarithm)103-5, 195,
219, 239-42; partial correlation 285, 311-12, 323; ladder of power 9,
193-4; simultaneous equation 100-2, 356-8; towards linearity
models 417-18, 450; spurious 148-59; semi-logarithmic 153-9; to
regressions 347-8, 355, 361, 363 stationarity 335-65
Tanzania 50-1; export growth 336, transformed data 9, 11, 102-5
337, 343, 352-5; farm size/household transformed explanatory variables 9,
size 119, 462; government 13-16
expenditure/revenue 128-37, 239-42, transformed variables 9, 11-12
462; household expenditure/food trend stationary process (TSP) 349-51,
expenditure 113, 159; informal 355, 360, 365
sector 303-5, 307-8, 312, 326; trends (regression with time-series
monetary growth 350-1, 352-5, data) 335-65
360-1; money supply 335-7, 461-2; Tukey, John W. 35, 37, 60, 79, 112,
prices 335-9, 341, 348, 350-5, 204, 306
357-61, 363, 461-2 Tukey's ladder of transformation 9,
Tanzanian Economic Surveys 462 100-2, 356-8
terms of trade 161, 234-5, 237-8, 462 two-stage least squares 420-1, 424,
test-test-test approach 34, 36 442-5, 447, 451
testing: for autocorrelation with two-tailed test 67, 68
lagged dependent variable 386;
causality (Granger and Sims) 425-7; unbiased estimator 53-8
for cointegration 399-406; diagnostic unconditional mean 138
(initial model specification and UNDP (Human Development Report)
estimation) 4-7; downwards 33-5; 458, 460
heteroscedasticity 256-64; linear unit roots: Durbin-Watson statistic as
combination between regression test for 386-7; testing for non-
coefficients 195-7; for non- stationarity 351-6, 402
stationarity (unit roots) 351-6, unitary elasticity 194-5
386-7; non-zero linear restrictions univariate analysis 7-9, 16; modelling
229-31; parameter stability 231-7; an average 44-74; skewness and
quadratic term 197-8; significance of outliers in data 75-107
R2 225-6; for stationarity 349-56; urban population (energy
unitary elasticity 194-5; zero consumption) 154-6
restrictions 219-29 urbanisation 187-91
thin tails 81, 82-3, 94, 100 urban wages (heteroscedasticity)
three-stage least squares 448, 450-1 253-6
496 Index
variable selection 163, 164, 208, World Development Report 77, 455,
219 460--1
variance inflation factors 216, 217 World Fertility Survey 98, 456
variances 73, 185-6; derivation of World Socioeconomic Data 16, 50, 88,
(AR(l) model) 391; stabilising 91, 104, 146, 150, 152, 154, 158, 173,
356--9 187, 231-2, 281, 287, 292, 294, 379,
442, 450, 460--1
wages 27-8; and age of workers World Tables (World Bank) 455-6,
(heteroscedasticity) 253-6, 261-6 457, 459, 460--1
weighted least squares 269, 270--7
Wheeler, E. F. 90 X (as stochastic variable) 123--4
White, Halbert 276--8 Xi (appropriate weight) 273-6
White, Howard 233, 277
White, K. J. 470, 472 Y (as stochastic variable) 123--4
White's test 256, 261, 262, 263--4, 267, Y values, confidence interval for
270, 276-8 predicted 122-3
white noise error 367 Yang, M. C. 234, 462
women: life expectancy 87-90, 91-3,
94; sex bias against 87-90 zero-sum property 76-7
workers ( age/wages) 253-6, 261-6 zero mean 129, 131, 134, 169, 401
Working, H. 158 zero restrictions (testing) 219-29
World Bank 5, 455-6, 457, 459, zone of indecision (Durbin-Watson
460-1 test) 386, 389

You might also like